[ { "title": "2204.11595v1.Ultrafast_racetrack_based_on_compensated_Co_Gd_based_synthetic_ferrimagnet_with_all_optical_switching.pdf", "content": "Ultrafast racetrack based on compensated Co/Gd-based\nsynthetic ferrimagnet with all-optical switching\nPingzhi Li1*, Thomas J. Kools1, Bert Koopmans1and Reinoud Lavrijsen1\n1*Department of Applied Physics, Eindhoven University of Technology, Eindhoven, 5612\nAZ, Netherlands.\n*Corresponding author(s). E-mail(s): p.li1@tue.nl;\nKeywords: synthetic ferrimagnet, angular momentum compensation, current-induced domain wall motion,\nall-optical switching\nSpin-orbitronics [1, 2] and single pulse\nall-optical switching (AOS) [3] of magne-\ntization are two major successes of the\nrapidly advancing \feld of nanomagnetism\nin recent years, with high potential for\nenabling novel, fast and energy-e\u000ecient\nmemory and logic platforms. Fast current-\ninduced domain wall motion (CIDWM) [4]\nand single shot AOS [5] have been individu-\nally demonstrated in di\u000berent ferrimagnetic\nalloys. However, the stringent requirement\nfor their composition control [5, 6] makes\nthese alloys challenging materials for wafer\nscale production [3]. Here, we simultane-\nously demonstrate fast CIDWM and energy\ne\u000ecient AOS in a synthetic ferrimagnetic\nsystem based on multilayered [Co/Gd] 2.\nWe \frstly show that AOS is present in\nits full composition range. We \fnd that\ncurrent-driven domain wall velocities over\n2000 m/s at room temperature, achieved by\ncompensating the total angular momentum\nthrough layer thickness tuning. Further-\nmore, analytical modeling of the CIDWM\nreveals that Joule heating needs to be\ntreated transiently to properly describe the\nCIDWM for our sub-ns current pulses. Our\nstudies establish [Co/Gd]-based syntheticferrimagnets to be a unique materials plat-\nform for domain wall devices with access to\nultrafast single pulse AOS.\nResearch in spintronics over the last decades\nhas demonstrated a possibility for further evo-\nlution of solid-state electronic application plat-\nforms beyond CMOS. The racetrack memory [7],\nan envisioned novel data storage device concept\nbased on using magnetic domain walls (DWs) [8],\nutilizes frontier spintronic mechanisms to func-\ntion as an ultra-dense on-chip memory with an\noperation speed comparable to low-level cache\n[9]. The e\u000eciency of racetrack memory relies\non the fast CIDWM in a material with per-\npendicular magnetic anisotropy. The maximum\nvelocity was signi\fcantly enhanced over the years\nby combining spin-orbit torques (SOTs) and the\nDzyaloshinskii-Moriya interaction (DMI) [10{12]\nwith synthetic antiferromagnets (SAFs), which\nresulted in reported DW velocities close to 750\nm/s [13].\nDespite the large improvement, the energy e\u000e-\nciency is still limited due to the weak strength\nof the antiferromagnetic (AF) coupling. There-\nfore, the materials platform of rare earth (RE)-\ntransition metal (TM) compounds garnered con-\nsiderable attention, promising faster CIDWM due\nto the much stronger direct AF coupling than the\n1arXiv:2204.11595v1 [cond-mat.mes-hall] 25 Apr 20222 Article Title\nindirect exchange coupling [14] utilized in SAFs.\nFurthermore, the SOTs used to drive the DW\npromise to be highly e\u000ecient in the RE-TM sys-\ntems as a result of the long spin coherence length\n[15]. Cosequently, high velocity CIDWM has been\nreported in Co-Gd-based ferrimagnetic alloy sys-\ntems [4, 16] when the angular momentum in the\nmagnetic material is compensated, being at least\na factor of three faster than that of the previously\nreported SAFs.\nBesides the e\u000ecient CIDWM, single pulse all-\noptical switching (AOS) of the magnetization\n[5] in the RE-TM systems has obtained signi\f-\ncant attention thanks to its sub-picosecond [17]\nenergy e\u000ecient [3, 18, 19] magnetization switch-\ning enabled by the ultrafast angular momentum\ntransfer upon laser excitation [20]. This can be\nuseful as a new generation of ultrafast magnetic\nmemory, as well as a data bu\u000ber between electron-\nics and integrated photonics [3, 21, 22]. Recently,\na synthetic ferrimagnetic system based on a Pt/-\nCo/Gd [19, 23] layered structure has shown high\nrobustness [24] for such a hybrid integration.\nThese kinds of synthetic ferrimagnets have\nsome distinct advantages over RE-TM alloys. For\ninstance, AOS is not limited by the exact compo-\nsition [25]. They also withstand thermal annealing\n[26] and o\u000ber easier magnetic composition control\nat wafer scale than the alloy system, as well as bet-\nter access to interface engineering. Therefore, it\nhas been proposed that such a materials platform\nhas high potential to realize a hybrid integration\nof DW memory in photonic platforms to further\nenhance their storage density [21, 27].\nSo far, the CIDWM of Co/Gd bilayers [27,\n28] has been investigated. However, the highest\nreported velocity, achieved at cryogenic conditions\n[28], was several times lower than that reported\nin alloys [4], in part due to large net angu-\nlar momentum, low compensation temperature as\nwell as DW pinning e\u000bects. In this report, we\ntherefore propose a materials platform based on\nthe [Co/Gd] 2synthetic ferrimagnet capable of\naccommodating both e\u000ecient CIDWM at room\ntemperature (RT) and single-pulse AOS, with\ncompatibility for wafer scale production as well as\nrobustness for engineering.\nThe magnetostatic and AOS properties of\nthe [Co/Gd] 2materials platform under investi-\ngation in this work are shown in Fig. 1. ATa/Pt/[Co/Gd] 2/TaN multilayer is deposited on\na Si/SiO 2substrate, where the \frst Gd layer is\nwedged (see Fig 1.a and Methods) to tune the\nnet magnetic moment. Due to the proximity e\u000bect\nby the Co, a net magnetization is induced in the\nGd which drops o\u000b steeply away from the Co\ninterface [29]. Compared to the bilayer Co/Gd\n[23, 28], we double the magnetic volume of the Co\nin [Co/Gd] 2, while tripling the number of inter-\nfaces at which magnetization is induced in the Gd.\nThis allows us to more easily compensate both\nthe angular momentum and magnetic moment of\nthe system at RT by varying the individual layer\nthicknesses.\nWe measured hysteresis loops by polar MOKE\nat RT across the thickness range of Gd between\n0-1.6 nm. In Fig. 1.b we plot the MOKE sig-\nnal intensity, de\fned as the di\u000berence in signal\nbetween positive and negative saturation, as well\nas the coercivity obtained from these hysteresis\nloops. The loops corresponding to the colored dots\nin Fig. 1.b are presented in Fig. 1.c.\nThe 100% remanence indicates a well-de\fned\nperpendicular magnetic anisotropy as well as tight\nexchange coupling between the layers. Impor-\ntantly, at the Gd thickness of 0.97 nm, the MOKE\nsignal switches sign (see black curve in Fig. 1.b).\nThis can be observed as well in Fig. 1.c as an\ninverted hysteresis loop. At this thickness the\nmagnetization is compensated, i.e. the Co and\nGd magnetization cancel each other. This is fur-\nther evidenced by the divergence in coercivity,\nand con\frmed by superconducting quantum inter-\nference device measurements (see Sup. A). Thus,\nupon increasing the thickness of Gd, the mag-\nnetic balance is shifted from being Co-dominated\nto Gd-dominated.\nWith respect to the magnetostatics, we note\nthat the magnetization and angular momentum\ncompensation thickness in these [Co/Gd]-based\nferrimagnets are similar but not identical due to\nthe di\u000berent Land\u0013 e g-factors of Co ( \u00182.2) and Gd\n(\u00182.0). Therefore, when discussing compensation\nwithout further speci\fcation in the remainder of\nthis text, we refer to angular momentum compen-\nsation as this is the relevant condition for e\u000ecient\nCIDWM [4, 28, 30].\nTo investigate whether our 4-layer Co/Gd sys-\ntem is all-optically switchable, and how it depends\non compensation conditions, we performed AOSArticle Title 3\na) b) c)\nCo(0.6)Gd(0-1.6)Co(0.6)Gd(1.5)TaN(4)\nPt(4)\nTa(4)d)\n1234567\nCo-dominated\nGd-dominated\nFig. 1 a): Schematic illustration of the layer stack used in our study (thickness in nm). b): Result of static MOKE study,\nwhere the MOKE signal (black) is de\fned by the di\u000berence in signal intensity between saturation at positive and negative\napplied \feld. Coercivity (red) was measured at a \fxed scanning speed of 10 mT/s (red). c): Hysteresis loops measured at the\nsample thicknesses marked by the coloured dots in b). d): Threshold \ruence for all-optical switching across the wedge shown\nin a). Inset shows toggle switched domains at two sides of the compensation thickness for varying amount of subsequent\npulses.\nexperiments by illuminating the wedge with sin-\ngle femtosecond laser pulses with varying pulse\nenergy, and examined the sample under a polar\nKerr microscope. We \fnd that single pulse AOS is\npresent at every thickness on the [Co/Gd] 2wedge.\nAn example of the resulting toggle switched\ndomains around the compensation boundary is\nshown in the inset of Fig. 1.d, which can be\nobserved from the contrast inversion (see Sup. B\nfor more images). This distinct feature that the\nAOS occurs both in the deep Co-dominated and\nGd-dominated composition is unlike that of alloys\n[18] and Co/Tb multilayers [31], in which AOS\nis only present within \u00061.5 % of the composition\nwindow. This shows the \rexibility of our system\nin terms of AOS engineering.\nWe further plot the threshold \ruence as a func-\ntion of Gd thickness, calculated from the switched\narea and the corresponding pulse energy [23],\nin Fig. 1.d. We \fnd that the threshold \ruence\ndepends signi\fcantly on the material composition,\nand is at a minimum around the compensation\npoint. Initially, the threshold \ruence decreases\nmonotonously with Gd thickness up to 1 nm, with\na reduction of the threshold \ruence of more than\n50% reaching 0.9 mJ/cm2, which corresponds to\n25 fJ for a 502nm2domain. Such a switching\nenergy is an order of magnitude lower than that\nof typical SOT and spin-transfer torque switching\nof a ferromagnet [32].\nAlthough a full dissection of the exact mech-\nanism behind the thickness-dependence of the\nthreshold \ruence is beyond the scope of this work,\nwe note that previous studies have shown thatreducing the Curie temperature of layered sys-\ntems can lead to a reduction of the threshold\n\ruence [6, 19, 23, 25]. In our case, as the Gd layer\nincreases, the Curie temperature of the total stack\nis expected to drop, because the mutual exchange\nstabilization of the two Co layers is weakened,\nleading to a reduction of threshold \ruence.\nWith compensation and AOS at RT con\frmed,\nin the remainder of this work we demonstrate\nand explain the DW velocities of over 2000 m/s,\nas achieved in [Co/Gd] 2. In the case of such an\nAF coupled system with DMI, two contributions\nmainly drive the coherent motion of the DW: the\nDMI-torque and the exchange coupling torque.\nThese two torques give rise to spin dynamics\nupon excitation by the spin Hall-e\u000bect-induced\nspin accumulation, and both lead to DW motion\nalong the current direction (for positive DMI and\nspin Hall angle [11, 33]).\nIt is well understood that the exchange cou-\npling torque only drives the DW motion e\u000eciently\nnear the compensation point [13, 28]. Adding\nadditional AF coupled layers to a ferromagnet\ncan facilitate this, however, can also increase the\nangular momentum of the DW to be translated.\nNonetheless, following the discussion in Sup. C, we\nstill expect a signi\fcant increase of the DW veloc-\nity in our study, as the enhancement of exchange\ncoupling torque due to compensation signi\fcantly\noverwhelms the e\u000bect of added magnetic inertia.\nFor this reason, we experimentally determined\nthe DW velocity from the CIDWM in [Co/Gd] 2\nas a function of lower Gd thickness. This study\nis performed in the same stack showcased in Fig.\n1.a. The wedged thin \flm was patterned into4 Article Title\na)\nb)c)\nd) e) f)\n0.4 ns\nFig. 2 a): Example image from DW motion experiments performed by di\u000berential polar Kerr microscopy, where the setup\nis illustrated schematically as a pulse generator and a 6 GHz bandwidth oscilloscope. The DW position after applying\nconsecutive pulses is shown. b): The measured DW velocity as a function of current density for three Gd layer thicknesses.\nc): The DW velocity as a function of Gd layer thickness for various current densities, where 10 % error bars are not shown.\nHere we marked the quanti\fed magnetization compensation thickness tMobtained from polar MOKE measurements as\nwell as the corresponding angular momentum compensation thickness tAbased on the estimation from magnetometry\nmeasurement. d): Compensation thickness as obtained from polar MOKE measurements of the [Co/Gd] 2at various ambient\ntemperatures. e): The temporal pro\fle of the current pulse used in the experiment (normalized by the peak current density),\nand its subsequent temperature rise (as simulated with COMSOL) at 3.6 TA/m2. f): Result of the theoretical 1-D modeling:\nthe DW velocity as a function of Gd thickness, where the magnetization pro\fle as well as its temperature dependence is in\nline with experimenes, while the current pulse and temperature pro\fle follow the pro\fle given in d). The velocity maximum\nis marked with a grey dot. The vertical dashed line indicates the angular momentum compensation thickness.\nmicro-structures as shown in Fig. 2.a (for details\nsee Methods and Sup. Fig. E4.b). The magnetic\ndomains were imaged by polar Kerr microscopy\n(see Fig. 2.a). Short current pulses of around 1 ns\nduration with varying amplitudes are injected into\nthe structures to determine the DW velocity (see\nSup. D). We observe that the DW moves coher-\nently along the current direction (see Fig. 2.a).\nThis con\frms a left-handed N\u0013 eel-type DW in our\nsystem as a result of a positive DMI, and a posi-\ntive spin Hall angle [11, 33], as is characteristic of\nthe Pt/Co interface [11].In Fig. 2.b we compare the resulting DW veloc-\nity as a function of peak current density for a\nCo-dominated (0.69 nm), compensated (0.93 nm)\nand Gd-dominated stack (1.28 nm), considering\nthe magnetic composition at RT. We demonstrate\nthat DW velocities of over 2000 m/s can be\nachieved at the highest current densities, at least\nthree times larger than that observed in SAFs [13].\nFurthermore, we notice from the crossing point\nbetween the red and black curve that the Gd\nthickness at which the highest velocity is observed\nbecomes larger with increasing current density. We\nattribute this to the e\u000bect of Joule heating inducedArticle Title 5\nby the current pulses. Earlier studies show that the\nGd magnetization is quenched more severely dur-\ning Joule heating than that of Co [28, 29, 34, 35].\nThe consequence is that at higher temperature rel-\natively more Gd than at RT is needed to achieve\ncompensation. This fact is con\frmed for our stacks\nby the upward shift of the compensation thickness\nwith increasing sample temperature using polar\nMOKE (see Fig. 2.d). We \fnd that this hypothesis\nof a Joule-heating-induced shift of the compensa-\ntion thickness explains the CIDWM experiments\nwell. To understand this, it useful to consider two\nregimes in Fig. 2.b in more detail.\nForJ < 2:3 TA/m2, before saturation e\u000bects\nstart to occur and for limited heating, we observe\na steady rise of the velocity, consistent with earlier\nwork on CIDWM in materials with AF coupling.\nAs expected, the compensated sample is driven\nmost e\u000eciently. We visualize this further by plot-\nting the DW velocity as a function of Gd thickness\nacross the Co and Gd-dominated regimes, as is\nshown in Fig. 2.c. Here, we observe that the DW\nvelocity in this range of current densities spikes\nat the RT compensation thickness, further prov-\ning the origin of the velocity increase is due to the\nenhanced exchange coupling torque.\nIn contrast, for J >2:3 TA/m2, the DW veloc-\nity saturates both for the Co-dominated and the\ncompensated sample. This is typical behaviour of\nthe CIDWM of a ferromagnet [11, 13, 28, 33, 36],\nwhich is limited by the DMI-torque. Contrar-\nily, the Gd-dominated sample shows a consistent\nvelocity rise typically associated with a compen-\nsated system [13]. This is consistent with the\nhypothesis that the increased Joule heating shifts\nthe compensation point to larger Gd thicknesses\nduring the pulse, and is supported by the velocity\npeak shift observed in Fig. 2.c.\nIn order to quantify the e\u000bect of the change\nin compensation thickness due to Joule heating\non the CIDWM, we \frst numerically investigate\nthe temperature transient for the current pulses\napplied in the experiment (see Sup. E). In Fig.\n2.e we plot a typical current density pro\fle and\nthe corresponding modelled temperature change\nin the device. We observe a delay of \u00180:4 ns\nbetween the peak current density, and the peak\ntemperature. This is a consequence of the factthat the pulse duration approaches the character-\nistic time scale of the heat transport between the\nsample and the substrate (See Sup.E).\nImportantly, around the peak current den-\nsity where the driving torque exerted on the DW\nis largest, the sample temperature is changing\nrapidly from 0 to 50 % of maximum temperature\nwithin 0.4 ns. Therefore, the net angular momen-\ntum changes rapidly during the application of\nthe pulse. This makes assigning an e\u000bective tem-\nperature during DW motion like previous studies\n[4, 28, 35, 37] inaccurate.\nTherefore, we incorporated the transient cur-\nrent density pro\fle, and the corresponding tem-\nperature rise into an analytical 1-dimensional\nmodel of the CIDWM (see Sup. F), where we\nintroduce a thickness and temperature dependent\nmagnetization pro\fle following our SQUID mea-\nsurements (see Sup. A). The resulting modelled\nDW velocity as a function of thickness and current\ndensity is plotted in Fig. 2.f.\nWe \fnd that the velocity peak shift with\nincreased current density as shown in Fig. 2.b is\nwell described by our model. Furthermore, we also\n\fnd that the rapidly changing angular momentum\nduring the current pulse gives rise to a broad-\nening of the velocity peak compared to the case\nwith constant temperature previously described in\n[4, 28](see Sup. G).\nThese results suggest that in general transient\ntreatment of the temperature pro\fle is neces-\nsary to properly address the DW dynamics for\nultrashort current pulses. Furthermore, to bene\ft\nfrom large exchange coupling torque at elevated\ntemperature, which is a likely scenario in high-\nspeed electronics, pushing the composition into\nthe Gd-dominated regime is required.\nIn summary, we have presented a materi-\nals platform of a synthetic ferrimagnet based\non [Co/Gd] 2which can be e\u000bectively tuned\nnear the compensation point via layer thickness,\nand exhibits e\u000ecient single pulse AOS. We also\nexperimentally and numerically showed that fast\nCIDWM with velocities over 2000 m/s can be\nachieved with the help of angular momentum\ncompensation. Finally, we addressed the transient\nJoule heating e\u000bect during CIDWM, which gives\nimportant insight into our experimental obser-\nvations and potential technological implementa-\ntion. Our results therefore lay the foundation\nfor an e\u000bective materials platform that exhibits6 Article Title\nboth ultrafast CIDWM and AOS, paving the way\nfor synthetic ferrimagnets to become the new\nparadigm of ferrimagnetic spintronics and pro-\nviding a jumping board for further integration\nbetween photonics and spintronics.\nMethods. Magnetic \flms were deposited on Si\nsubstrates with a 100 nm thermally oxidized SiO 2\nlayer by D.C. magnetron sputtering in a system\nwith a base pressure of \u00185\u000210\u00009mBar. A\nthickness wedge is created with the help of a\nmoving wedge shutter during sputtering. From\nthese thin \flms, nanostrips were fabricated using\nelectron-beam lithography and lift-o\u000b. The gold\ncontacts were made by wire bonding. The out-\nof-plane component of the magnetization (M z)\nof the nanostrips was measured by polar Kerr\nmicroscopy. An in-plane \feld of 200 mT along\nthe current direction, combined with a 3 TA/m2\ncurrent pulse was used to nucleate domains.\nBoth MOKE and Kerr Microscopy characteriza-\ntion were performed using a 700 nm continuous\nwave laser in the polar con\fguration. Here, we\nchose the wavelength of the laser light (around 700\nnm) such that the detected signal is only sensitive\nto the magneto-optical response from the Co layer\nand Kerr rotation is maximized.\nSupplementary information. All supplemen-\ntary information except for digital \fles are pre-\nsented in the Appendix.\nContribution. P.L. and T.J.K. contributed\nequally to this work in terms of design and con-\nduct of the project. B.K. and R.L. supervised the\nproject. All authors contributed to the writing of\nthe manuscript.\nAcknowledgments. This project has received\nfunding from the European Union's Horizon\n2020 research and innovation programme under\nthe Marie Sklodowska-Curie grant agreement\nNo.860060. This work is also part of the Gravita-\ntion programme `Research Centre for Integrated\nNanophotonics', which is \fnanced by the Nether-\nlands Organisation for Scienti\fc Research (NWO).\nCompeting Interests. Authors declare no\ncon\ricts of interest.Appendix A SQUID\ncharacterization\nIn order to characterize the magne-\ntostatic properties of the sample, a\nseries of 6 samples of composition\nTaN(4)/Pt(4)/Co(0.6)/Gd( tGd)/Co(0.7)/Gd(1.5)\n/TaN(4), with tGd= [0:2;0:4;::;1:2;1:4] nm were\ngrown. Using VSM-SQUID, out-of-plane hystere-\nsis loops were measured, which are shown in Fig.\nA1.a. The decreasing moment when going from\na middle Gd thickness of 0.4 to 1.4 nm, and the\ndiverging coercivity at the minimum of the mea-\nsured moment are indicative of a compensated\nsystem. This can also be seen in the plot of the\nmagnetic moment normalized by the sample area\nin Fig. A1.b. The temperature dependence of the\nmagnetization was also investigated, and a typical\nresult is shown in Fig. A1.c. In order to estimate\nthe Curie temperature of the stack and obtain an\nexpression for the magnetization Mtot, we follow\nearlier work on 3d-4f ferrimagnetic alloys in \ft-\nting a critical exponent to the data of the form\n[30, 38] given by:\nMtot=MCo(T)\u0000MGd(T) =\nMCo0\u0012\n1\u0000T\nTC\u0013\u0018Co\n\u0000MGd0\u0012\n1\u0000T\nTC\u0013\u0018Gd\n;\n(A1)\nwhereMGd0andMCo0and\u0018Gd,\u0018Coare the\nmagnetization at zero temperature and the critical\nexponent of Co and Gd, respectively. We assume\nMCo0= 1:4 MA/m, and \fnd that the critical\nexponents\u0018Co= 0.50,\u0018Gd= 0.72, and TC= 450 K\ndescribe the temperature dependence of the mag-\nnetic moment well (Fig. A1.c). We note that the\nmagnetization in the Gd is an ill de\fned quantity,\nsince the Gd transitions from partially magnetized\ndue to the proximity e\u000bect, to truly ferromag-\nnetic at low temperature. We therefore estimate\nthe value of MGd0by considering the known thick-\nness dependence of the magnetic moment in the\nstackmtot. This quantity is calculated as the sum\nof that of the individual layers as follows:\nmtot=mCo1+mGd1+mCo2+mGd2\n=MCo(T) (tCo1+tCo2)\u0000Article Title 7\na) b) c)\nFig. A1 VSM-SQUID characterization of the out-of-plane moment of the Co(0.6)/Gd(x)/Co(0.7)/Gd(1.5) material system.\na): Hysteresis loops across the compensation point at room temperature. b): Room temperature magnetic moment per unit\narea as a function of middle Gd thickness. c): Typical temperature dependence of the magnetic moment in a Co-dominated\nquadlayer sample. The red line indicates the best \ft of equation A1.\nMGd(T) (2tGd1+tGd2);(A2)\nwhere:\nmCo1=MCo(T)tCo1\nmCo2=MCo(T)tCo2\nmGd1= 2MGd(T)tGd1\nmCo1=MGd(T)tGd2: (A3)\nHere,tCo1,tCo2,tGd1,tGd2are the thicknesses of\nthe bottom (coming from the substrate) and top\nCo and Gd layers, respectively. We \fnd the red-\ncurve in Fig. A1 b, which represents equation (A2)\nfor the stack dimensions used in the experiment,\nwhich follows the SQUID data reasonably well by\nchoosingMGd0= 0.66 MA/m. It should be noted\nhere that this is about a factor of 2 lower than\nvalues typically found in CoFeGd alloys [30]. We\nargue that this is a consequence of the inhomoge-\nneous proximity induced magnetization in the Gd,\nwhich decays when moving away from the Co/Gd\ninterface, and is therefore concentrated at the\ninterface. In order to describe the magnetization\nfully, an intermixing pro\fle and proximity-induced\nmagnetization pro\fle needs to be assumed and\nsubstantiated, which is beyond the scope of this\nwork. We nonetheless choose this simple approach\nof homogeneous magnetization in the Gd, as the\nintra-layer exchange present in the Gd is expected\nto be strong enough to have the total angular\nmomentum in the Gd layer respond coherently to\nthe excitation of the DW, making the total angu-\nlar momentum present in the DW the relevant\nparameter and not the exact way it is distributed.Appendix B All optical\nswitching in\n[Co/Gd] 2\nIn addition to the meta-data of all-optical switch-\ning as shown in the main text (Fig. 1.d, repeated\nin Fig. B2.b), in Fig. B2.a we present a typical\nexcerpt of the polar Kerr microscope images used\nto obtain the threshold \ruence as a function of\nGd-thickness.\nAppendix C Velocity\nenhancement\nexpectation\nTo argue why a signi\fcant velocity enhancement is\nexpected regardless of the added magnetic volume\nfrom the multiple magnetic layers, it is instructive\nto discuss the ratio between the expected DMI-\ntorque-driven DW velocity in an uncompensated\nsamplevucto the exchange coupling torque-driven\nDW velocity in a compensated sample vc. Here we\ndiscuss this using an analytical approach based on\na 1-D model of DW motion [13, 28, 39] (see also\nsection F), the ratio is given by:\nvc\nvuc=~J\u0012SH\u0001\n2e\u000bDAuc\nAc: (C4)\nHere, we consider the fact that vuc, is limited\nby the strength of the DMI, following [4, 36]:\nvuc=\u0019D\n2P\nnjAnj;=\u0019D\n2Auc; (C5)8 Article Title\na)\nb)\n200 μm\n1.50\n1.30 \n1.05 \n0.96 \n0.80 \n0.50 \n0.30 \n0.10 Gd thickness (nm)\n180 160 140 120 100 80 60 40\nPulse enery (nJ)\nFig. B2 (a) Di\u000berential Kerr images (background taken\nat a Gd thickness of 0.1 nm) taken at di\u000berent locations\non the Gd wedge sample, which was illuminated by single\nlaser pulses with varying pulse energy. The inverted back-\nground contrast around 1 nm is due to the transition from\nCo-dominated to Gd-dominated. b): Calculated threshold\n\ruence as a function of Gd layer thickness. The threshold\n\ruence below and above compensation is calculated with\nthe average of 4 neighboring rows.\nwhereDis the surface DMI energy density, and A n\nandAucare the total angular momentum of the\nnthlayer and the full stack, respectively. On the\nother hand, vcis for strong enough exchange cou-\npling limited by the amount of angular momentum\ntransferred to the DW via the spin Hall-e\u000bect, and\ncan for a multilayered system be expressed as:\nvc=~J\u0012SHE\u0001\n4e\u000bP\nnjAnj=~J\u0012SHE\u0001\n4e\u000bA c:; (C6)where ~is the reduced Planck constant, Jis the\ncurrent density, \u0012SHE is the spin Hall angle for\nanti-damping-lik torque, eis the electron charge, \u000b\nis the Gilbert damping parameter, Acis the total\nangular momentum of the full stack, and \u0001 is the\nDW width, which shows no dependence on the\nDMI.\nTo estimate the expected velocity enhance-\nment, we consider equation (C4) for the simple\ncase where Auc=Ac= 1=2. This corresponds to\nthe situation where one antiferromagnetically cou-\npled layer is added to a magnetic system, which\nexactly compensates the angular momentum in\nthe original system. Within this simple model,\nand using equation (C4), the condition for DW\nvelocity enhancement then becomes:\nvc\nvuc=~J\u0012SH\u0001\n4e\u000bD>1: (C7)\nWe \fnd that for the parameters used in the\nsimulations (see Sup. F), a net gain in DW veloc-\nity can already be obtained at a current density\nof 0.28 TA/m2. Therefore, the exchange cou-\npling torque in our system can be expected to\nresult in a much higher velocity compared with\nthe uncompensated case where the DMI-torque\ndominates regardless of the inevitable increase in\nmagnetic volume for the synthetic ferrimagnets\nunder discussion.\nAppendix D Domain wall\nvelocity charac-\nterization\nIn this section, we discuss the experimental details\nof the characterization of the current-induced DW\nvelocity.\nWe \frst discuss the principle of the measure-\nment. To magnetically saturate the sample, we\n\frst reverse the magnetization in the wire with\nrespect to the rest of the device by a combined\naction of an intense current pulse and an in-\nplane \feld along the current direction. Then, N\ncurrent pulses with amplitude J, and pulse dura-\ntion \u0001t, displace the DW by a distance \u0001 L,\nwhich is recorded by polar Kerr microscopy. The\nresulting DW velocity, vDW, is then calculated as\nvDW=\u0001L\nN\u0001t. We applied multiple pulses of the\nsame amplitude to allow the DW to travel enough\ndistance from the start to the end of the wire toArticle Title 9\nminimize the errors of the length measurement\nand statistical errors caused by variation between\npulses as a result of wire edge roughness and other\npotential fabrication imperfections. For the case of\nhigh current density or high velocity, this process\nis repeated several times for better averaging due\nto lowNbeing needed to translate the DW along\nthe full wire length.\nThe method mentioned above requires the\nideal case of a current pulse with a square pro-\n\fle. Next, we discuss the pulse pro\fle used in our\nexperiment, based on which we de\fne our e\u000bec-\ntive pulse width and amplitude. The amplitude\nand duration of the pulse needs to be calibrated\ncarefully in order to clearly de\fne the DW veloc-\nity. In our measurement set-up, we terminate\nthe device with a 6 GHz bandwidth oscilloscope\n(characteristic impedance 50 \n), allowing us to\ncharacterize the electrical pulse pro\fle directly\nduring the measurements. The resulting voltage\nversus time pro\fle obtained from the oscilloscope,\nis then converted to current density values versus\ntime based on the DC resistance and the geom-\netry of the devices. Here, we have to note that\nin our calculations, only the thickness of Pt was\nconsidered in our calculation (assuming no cur-\nrent \rowing in both the ferromagnetic layers and\nTa). Such an assumption is made because Pt is\nknown to be more conductive than the other met-\nals used in the stack [28]. Therefore, the energy\ne\u000eciency of CIDWM in our study is a conservative\nestimate, since the current is likely also \rowing\nin the other metallic layers. To avoid dramatic\nheat e\u000bects induced by long current pulses, which\nmight screen much of the physics, we used ultra-\nshort pulses down to 1 ns in our study. The pulse\namplitude can reach up to 4 \u00021012A/m2without\nfully demagnetizing the sample. A typical mea-\nsurement of the current pulse pro\fle is shown in\nFig. D3.a. In our measurement, we de\fne the pulse\namplitude to be the peak value (1.8 \u00021012A/m2in\nthis case). The pulse exhibits a spike shape, which\nrequires us to de\fne an e\u000bective pulse width. We\nwe de\fne the e\u000bective pulse width in our mea-\nsurement by taking the full width half maximum\n(0.8 ns) of the active area (see gray region in\nFig. D3.a) of the pulse obtained by subtracting\nthe pulse shape by a strong, experimentally deter-\nmined, pinning threshold (0.5 TA/m2). Thus, the\ntime needed for the current pulse to pull up andthe decay of the pulse below the pinning thresh-\nold is not considered, since in this region the DW\nis not displaced or to a limited extent. To de\fne\nthe pinning threshold, we applied over 500 current\npulses when the DW is at various positions of the\nwire. The pinning threshold is then de\fned as the\ncurrent density where no detectable movement of\nthe DW is found. A further reason for using short\npulses is the fact that the temperature rise during\npulse time can be easily suppressed by the heat\ndissipation and capacity of the substrate since\nthe heat front is still in the propagation regime.\nThis also leads to a transient delay of temperature\nrise (see the next sections for its in\ruence on the\nCIDWM).\nTo justify the e\u000bective pulse parameters,\nwe compare the measurement result with that\nobtained from a longer pulse (L) and a middle\nlong pulse (M) shown in Fig. D3.b, of which the\nshape is closer to a square pulse. As seen from\nFig. D3.b, the L and M pulses both have an over-\nshoot and a slanted decay at the end of the pulse.\nWe compensate for this e\u000bect by subtracting the\nDW displacement induced by a shorter pulse from\nthat of a longer pulse. Since the di\u000berence between\nthese pulse shapes resembles a square pulse shape\nwith known pulse amplitude and width, the DW\nvelocity can be more accurately calculated. How-\never, in order to neglect the contribution from\nthe part of the pulse that was not subtracted, it\nrequires the thermal e\u000bect to be low (since the ini-\ntial spike contributes heat as well). Thus, it is only\napplicable for low current density.\nNow, we introduce a reference pulse (R)\n(shown in red), which resemble the spike of L and\nM, such that its di\u000berence with L and M will\nresults in a square pulse with a duration longer\nthan 1 ns as shown in Fig. D3.b. We calculated the\nDW velocity for the samples used in our experi-\nment (shown in Fig. 2) based on the pulse time\ndi\u000berence between L and M (L-M), L and R (L-R),\nas well as M and R (M-R), which are plotted in\nFig. D3.c (current amplitude used as shown in Fig.\nD3.b). We compare the obtained results with that\ncalculated based on the short pulse (S) used in\nthe experiment shown in the main paper (see blue\ncurve in Fig. D3. b), which di\u000bers in amplitude\nfrom the pulse pro\fle in Fig. D3. a.\nWe observe that the DW velocities using these\napproaches are in the same order of magnitude,\nfollowing the same general dependence on Gd10 Article Title\na) b) c)\nFig. D3 a): The pulse pro\fle of a typical pulse used in the DW velocity measurements. The strong pinning region which\nwas excluded during the pulse width estimation was marked as brown, the active region from where the e\u000bective pulse width\nis obtained is mark as gray. The pulse width and amplitude obtained for this characteristic pulse pro\fle are marked in the\nplot. b): The pulse pro\fle of a short pulse (S) used for the experimental results shown in the main text, a middle long pulse\n(M), a long pulse (L) and a reference pulse (R). The estimated time di\u000berence between those pulses are given and will be\nused to calculate the DW velocity. c): The DW velocity obtained from a short pulse, and other di\u000berential pulse method.\nthickness. The DW velocity calculated using S is\nfound to be systematically lower. Here we choose\nto be conservative instead of trying to compensate\nfor this discrepancy, as \fnite thermal activation\nis expected to increase the DW velocity by reduc-\ning the DW friction [11, 28, 40], as is the case for\nlong pulses. So far, little study has been spent on\nsuch an issue in the \row regime of the CIDWM,\nalthough the absolute magnitude of enhancement\nwas once shown to be 20 % for a SAF [41] for 40 K\nof temperature increase. So we make a compromise\nby taking a large but safe margin avoiding over-\nestimation of the DW velocity as well as possible\nover-claim (for large velocity).\nAppendix E COMSOL\nsimulations of\nJoule heating\nIn this section, we present our study on the tem-\nperature rise during the application of the current\npulses. This is a crucial ingredient for the anal-\nysis of the experimentally observed DW motion,\nas well as design considerations for future appli-\ncations. Here, we based our study on COMSOL\nsimulations, which employs multiphysics modeling\nincluding thermal transport and electrical con-\nduction. In our model, we de\fned the geometry\nbased on the real physical dimensions of the device\nused in this study (see Fig. E4.b) and the mate-\nrial parameters either from the COMSOL material\nlibrary or from our measurements (see Table E1),\nwhich takes into account temperature-dependente\u000bects. We model the metallic stack as a dou-\nble layer consisting of a 8 nm non-conductive\nlayer, and a 6 nm conductive layer, of which\nthe conductivity is obtained from RT DC mea-\nsurements, while the density, heat capacity and\nthermal conductivity of the two layers are kept the\nsame.\nTable E1 Parameters used in COMSOL\nsimulation for Joule heating induced by the\ncurrent pulse.\nParameter @ RT Value Unit\nThickness of the Si substrate 0.5 mm\nThickness of SiO 2 100 nm\nDiameter of wire bond 50 \u0016m\nHeat Capacity of Si Substrate 678 J/(kg \u0001K)\nDensity of Si 2320 kg/m3\nThermal Conductivity of Si 134 W/(m \u0001K)\nHeat Capacity of SiO 2Substrate 730 J/(kg \u0001K)\nDensity of SiO 2 2200 kg/m3\nThermal Conductivity of SiO 2 1.4 W/(m \u0001K)\nThickness of conductive metal \flm 6 nm\nThickness of non-conductive metal \flm 8 nm\nElectrical Conductivity of Metal 0.893 106\u0001S/m\nHeat Capacity of the Metal 133 J/(kg \u0001K)\nDensity of the metal 21450 kg/m3\nThermal Conductivity of the Metal 71.6 W/(m \u0001K)\nRelative permittivity 10000 1\nMinimum Mesh Dimension 0.5 nmArticle Title 11\nElectrical stimuli are applied through a circu-\nlar region (yellow disks in Fig. E4.b), of which\nthe size is close to the physical size of the electri-\ncal contacting wire bonds in our experiments. The\nwire is signi\fcantly heated compared with the rest\nof the metal structure owing to its large current\ndensity. Air convection is present only at the top\nsurface to emulate the experimental conditions.\nFollowing these notions, we computed the temper-\nature rise relative to RT as a result of a current\npulse similar to the one used in the experiment,\nas indicated in Fig. E4.a. In order to describe\nthis pulse pro\fle, we \ft an empirical function J(t)\nconsisting of the sum of three Gaussians:\nJ(t) =3X\ni=1Ai\nwip\n\u0019=2exp \n\u00002(t\u0000tc,i)2\nw2\ni!\n:(E8)\nThe best-\ft parameters are summarized in\ntable F3. The same pro\fle is used to de\fne\nthe voltage pulse that is applied to the system\nin COMSOL. The resulting temperature rise for\nvarious peak current densities is plotted in Fig.\nE4.c. One important observation here, is that the\nheating of the magnetic layer is delayed by approx-\nimately 0.4 ns with respect to the current pulse.\nConsequently, when the SOT is maximized at the\npeak of the current pulse, the angular momen-\ntum is still rapidly changing. Since the exchange\ncoupling torque e\u000eciency is linked to the angu-\nlar momentum balance between the Co and Gd,\nthis transient heating leads to a transient DW\nmobility.\nFor Joule heating, the expected temperature\nrise depends quadratically on the current den-\nsity. In order to describe the temperature rise for\nany value of the peak current density J0, we \ft\nan asymmetric sigmoidial function to the tem-\nperature rise at various current densities given\nby:\nT(t) =y0+b01\u0000exp\u0010\nt\u0000tc\nl2\u0011\n1 + exp\u0010\n\u0000t\u0000tc\nl1\u0011: (E9)\nThe resulting \fts for various current densities are\nshown in Fig. E4.d. During this \ft all parameters\nexcept the amplitude of the peak b0were taken\nconstant. The found amplitude, which describesthe heating is then plotted as a function of cur-\nrent density in Fig. E4.e. We \ft a simple quadratic\nfunction to this quantity:\nb0(J0) =k0J2\n0; (E10)\nwhich is shown as the red line. The correspondence\nbetween the \ft and the temperature rise modelled\nby COMSOL suggests that Joule heating plays the\ndominant role in our system.\nIn summary, based on the best \ft to COMSOL\nsimulations and the quadratic dependence of the\ntemperature change on current density, we made\nan estimation of the temporal pro\fle of the tem-\nperature rise in our sample. This asymmetric sig-\nmoidal pro\fle with the quadratic prefactor given\nin equation E9 and E10, will be used to dynam-\nically describe the change in temperature in the\nnumerical 1-dimensional simulations of CIDWM\nin our [Co/Gd] 2samples.\nTable E2 Summary of the \ftting parameters\nin equations E8 and E9, found to best describe\nthe experimental current pro\fle and the\nmodeled temperature rise due to Joule\nheating, respectively.\nParameter Value Unit\ntc,1 -2.85 ns\nw1 0.42 ns\nA1 0.54 TA m\u00002ns\ntc,2 -3.29 ns\nw2 1.23 ns\nA2 0.29 TA m\u00002ns\ntc,3 -2.29 ns\nw3 0.59 ns\nA3 0.31 TA m\u00002ns\ny0 293.15 K\ntc 2.15 ns\nl1 0.11 ns\nl2 1.94 ns\nk0 20.7 K m4TA\u0000212 Article Title\na) b)\nc) e)\nSiO2 (100 nm)\nSi (500 μm)200 μm 600 μm 100 μm \n50 μm 50 μm 2 μm \n10 nm \nd)\nFig. E4 Joule heating analysis and current pulse de\fnition for the model. a): Fit of three Gaussians (eq. E8) to a typical\nexperimental current pulse pro\fle. b): Schematic of the device geometry used in the experiment, as well as in COMSOL. c):\nTransient temperature modeled using COMSOL for the experimental current pulse pro\fle at various peak current densities.\nd): Fits of equation E9 to the temperature pro\fle at various peak current densities J0. e): Parabola \ft to the prefactor b0\nin eq. E9 obtained from the analysis from d).\nAppendix F 1-dimensional\nmodel of\ncurrent induced\ndomain wall\nmotion\nIn this section the analytical 1D-model for\ncurrent-induced DW motion in the [Co/Gd] 2mag-\nnetic system is described. This model is an exten-\nsion to earlier work by Bl asing et al. [28], who\nintroduced the model to describe the DW dynam-\nics of a Co/Gd bilayer. We will \frst summarize\nthe main points of this model, noting that a full\ndiscussion, can be found elsewhere [41].\nThe model discussed below describes the\ndynamics of an up-down DW in a nanowire device\n[length (x-direction)\u001dwidth (y-direction)\u001d\nheight (z-direction)]. The orientation of the spins\nin the DW will be discussed in terms of spherical\ncoordinates \u0012and\u001e, which describe the polar and\nazimuthal angles of spins in the system, respec-\ntively. We de\fne \u0012= 0 and\u0012=\u0019to be the\npositive and negative out-of-plane ( z) direction\nrespectively, and \u001e= 0 and\u001e=\u0019to be pointingto the positive and negative x-direction, respec-\ntively. Within the DW, the spins rotate gradually\nfrom\u0012= 0 to\u0012=\u0019. In equilbrium, the DW angle,\nwhich we de\fne as the azimuthal angle of the spin\nin the middle of the DW will point along \u001e=\u0019\n(-x) due to the counterclockwise DMI.\nIn order to then derive the equations of\nmotion of the spins in the DW in the presence\nof non-conservative forces, like Gilbert damping\nand current-induced torques like the spin transfer\ntorque (STT) and spin Hall e\u000bect (SHE) torque,\nthe Rayleigh-Lagrange equation is solved:\n@L\n@pi\u0000d\ndt@L\n@_pi\u0000@F\n@_pi: (F11)\nHerepirefers to the free variables in the system,\nLandFdescribe the Lagrangian and the dissi-\npation function, respectively. Within our model\nof [Co/Gd] 2there are \fve free variables: The\nazimuthal angles \u001efor the separate layers, and the\nDW position q, from which the DW velocity _ qis\nderived:\n_q=\u0000\u0001\nsin\u0012_\u0012: (F12)\nHere it is assumed that the DW positions are\ntightly coupled, such that qis always the same forArticle Title 13\nthe di\u000berent layers. The quantities LandFare\nde\fned as:\nL=Z1\n\u00001!DW\u0000Tdx; (F13)\nand\nF=Z1\n\u00001P\u000b+PSTT+PSHEdx; (F14)\nrespectively, where P\u000b,PSTTandPSHE, are the\ndissipation functions for Gilbert Damping, STT\nand SHE respectively, Tis the \"kinetic energy\"\nof the DW which we take identical to Blaesing et\nal. [41], and \fnally !DWis the energy density of\nthe DW. In order to de\fne !DW,P\u000b,PSTTand\nPSHE, it is useful to introduce some compact nota-\ntion for parameters corresponding to a particular\nmagnetic layer. We will refer to parameters corre-\nsponding to speci\fc magnetic layers through the\nsubscripts 1, 2, 3 and 4, where the numbering\nrefers to the layer in the Pt/[Co/Gd] 2structure\nstarting with layer 1 from the bottom upwards\n(i.e. layer 1 is the Co layer interfaced with the Pt\netc.). Using this notation !DWis given by:\n!DW=D1\n\u0001cos\u001e1sin\u0012\n+4X\ni=1Kitisin2\u0012+4X\ni=1Aex,iti\n\u00012sin2\u0012\n+3X\ni=1Jex\u0000\ncos2\u0012\u0000cos (\u001ei\u0000\u001ei+1) sin2\u0012\u0001\n:(F15)\nHere,Dis the DMI constant, \u0001 is the DW width,\nKiis the anistropy energy density, tiis again the\nthickness,Aexis the exchange sti\u000bness, and Jex\nis the exchange coupling strength. The DMI is\nassumed to only interact with the bottom Co layer\n(i= 1) as it originates from the Co/Pt inter-\nface, and no DMI at the Co/Gd interface has yet\nbeen reported. The \fnal term, describing the anti-\nferromagnetic exchange coupling has been chosen\nsuch that negative Jexpromotes antiferromagnetic\ncoupling.\nNext, the dissipation functions are de\fned as:P\u000b+PSTT+PSHE=4X\ni=1\u0000\u000bi;mi\u0001\n\ri\u0010\n_\u001ei2+ _q2=\u00012\u0011\n+4X\ni=1\u001diuimi\n\u0001\risin2\u0012_\u001ei\u0000\fiuimi\n\u00012\risin2\u0012_q\n\u0000~\u0012SHJ\n2esin\u0012\u0012cos\u001e1_q\n\u0001+ sin\u001e1cos\u0012_\u001e1\u0013\n;\n(F16)\nwhere\u000bis the Gilbert damping coe\u000ecient, \ris the\ngyromagnetic ratio, \fis the parameter describ-\ning the non-adiabatic component to the STT, ~is\nthe reduced Planck constant, \u0012SHis the spin-Hall\nangle,Jis the current density, eis the electron\ncharge, and \u001diis a variable that is equal to 1 and\n-1 in a Co layer ( i= 1;3) and Gd layer ( i= 2;4)\nrespectively. Finally, uiis given by:\nui=~Pi\ridi\n2emiJ; (F17)\nwherePiis the degree of polarization of the\ncharge current. It should be noted that the spin-\nHall current is in this work modelled to only\ninteract with the Co layer interfaced with the Pt.\nRecent work has indicated that the spin coherence\nlength in ferrimagnetic multilayers can be larger\nthan the thickness of the individual layers in this\nstudy [15]. Hence, it is possible that part of the\nspin current injected along the z-direction can also\ninteract with the other magnetic layers when it is\n(partially) transmitted through the \frst Co layer.\nIn order to then solve equation (F11) for the\n\fve degrees of freedom in our system, we evalu-\nate the integrals in equations (F13) and (F14) by\nassuming a typical Bloch pro\fle of the polar angle\n\u0012as a function of x:\n\u0012(x) = 2 tan\u00001\u0012\nexp\u0014x\u0000q(t)\n\u0001\u0015\u0013\n: (F18)\nThe resulting equations of motion in absence of an\napplied \feld for q,\u001e1,\u001e2,\u001e3,\u001e4are then \fnally\ngiven by:\n_q=\u0001\u0010P4\ni=1mi\u000bi\n\ri\u0001\u0011\u0012\n\u00004X\ni=1miui\fi\n\ri\u0001\n\u0000\u0019~\u0012SHJ\n4ecos\u001e1\u00004X\ni=1\u001dimi_\u001ei\n\ri\u0013\n(F19)14 Article Title\n_\u001e1=u1\n\u000b1\u0001+D1\u0019\r1sin\u001e1\n2m1\u000b1\u0001\n\u0000Jex\r1sin\u001e1\u0000\u001e2\nm1\u000b1+_q\n\u000b1\u0001(F20)\n_\u001e2=\u0000u2\n\u000b2\u0001+Jex\r2sin\u001e2\u0000\u001e3\n\u000b2m2\n\u0000Jex\r2sin\u001e1\u0000\u001e2\n\u000b2m2+_q\n\u000b2\u0001(F21)\n_\u001e3=u3\n\u000b3\u0001\u0000Jex\r3sin\u001e3\u0000\u001e2\nm3\u000b3\n\u0000Jex\r3sin\u001e4\u0000\u001e3\n\u000b3m3+_q\n\u000b3\u0001(F22)\n_\u001e4=u4\n\u000b4\u0001\u0000Jex\r4sin\u001e4\u0000\u001e3\n\u000b4m4+_q\n\u000b4\u0001:(F23)\nWe extended the model described earlier by\nBlaesing in three ways. First, we implement\nthe thickness and temperature-dependence of the\nmagnetization based on the SQUID measurements\ndiscussed in section A, in order to account for the\ndi\u000berent thicknesses of the samples investigated in\nthe experiment. Second, the temperature of the\nsample, contrary to earlier work, is also taken as\na dynamic parameter since we \fnd that for the\nultrashort current pulses investigated in the main\ntext, the sample is still heating up rapidly at the\npeak of the current pulse, when the driving torque\nis largest. Finally, we trivially extended the mod-\neled magnetic system from two to four coupled\nmacrospins.\nIn order to arrive at the results presented in the\nmain text, we numerically solve these \fve coupled\ndi\u000berential equations. For this we use the tempo-\nral current density pro\fle Jas given in equation\nE8 and table E2, taking into account the exper-\nimentally observed threshold current density of\n0.5 TA/m2(see Sup. D). The magnetization and\nits temperature dependence are implemented as\ndescribed in section A. All other physical parame-\nters used in the model are summarized in table F3.\nIn order to \fnally characterize the velocity fromthe model, we divide the DW displacement by the\nsame e\u000bective pulse width of 0.8 ns used to cal-\nculate the experimental domain wall velocity (see\nSup. D).\nWe \fnally note that for the realistic mate-\nrial parameters obtained from literature, the 1-D\nmodel generally predicts lower DW velocities than\nthe experiments. However, an increase of the DW\nwidth from 8 to 12 nm (both reasonable for\nthese kind of ferrimagnetic systems [47, 48]) could\nalready address this discrepancy in velocity with-\nout impacting the qualitative dependence of the\nDW velocity on the Gd thickness and current\ndensity.\nAppendix G Velocity peak\nbroadening\ndescribed by\ndynamic\ntemperature\nTo further substantiate the claim that dynamic\nsimulation of the temperature is important to\nproperly interpret the CIDWM experiments in our\n[Co/Gd] 2multilayers, we make a more elaborate\ncomparison between simulation with a dynamic\ntemperature pro\fle, and a constant temperature.\nFor this we use the same simulations as pre-\nsented in other parts of this work, and the same\ntemperature pro\fle for the dynamic temperature\nsimulations. For the constant temperature simu-\nlations we take the temperature to be 85% of the\nmaximum temperature of the dynamic tempera-\nture pro\fle, as this gives rise to a similar shift of\nthe thickness at which the peak velocity occurs.\nA typical comparison of the modelled DW veloc-\nity between static and dynamic temperature can\nbe seen in Fig. G5 for peak current densities J0of\n1.6, 2.6 and 3.6 TA/m2. It can be seen that for low\ncurrent densities the resulting velocity pro\fles are\nalmost identical, whereas for high current densi-\nties the velocity is more sharply peaked around the\ncompensation thickness. The main e\u000bect at high\ncurrent density of the dynamic temperature pro-\n\fle is a simultaneous lowering of the peak velocity\nand a broadening of the velocity peak, both of\nwhich can be understood from the magnetic com-\nposition changing rapidly throughout the peak ofArticle Title 15\nTable F3 Parameters used in 1D-modeling of CIDWM in [Co/Gd] 2.\nParameter Unit Co 1 Co2 Gd1 Gd2\ng - 2.2(a)2.2(a)2.0(b)2.0(b)\n\r 1011rad/s/T 1.93(c)1.93(c)1.76(c)1.76(c)\n\u000b - 0.1(d)0.1(d)0.1(e)0.1(e)\n\u0001 nm 8(f)8(f)8(f)8(f)\nJex mJ/m2-0.9(g)-0.9(g)-0.9(g)-0.9(g)\n\f - -0.5(h)-0.5(h)-0.5(h)-0.5(h)\nP - 0.1(i)0.1(i)0.1(i)0.1(i)\n\u0012SH - 0.08(j)0(k)0(k)0(k)\nD pJ/m 0.25(l)0(m)0(m)0(m)\naFrom [42, 43].\nbFrom [44].\ncCalculated using the layer's respective Land\u0013 e g-factor as \r=g\u0016B\n~.\ndFrom [11].\neAssumed similar to that of Co [45].\nfSimilar to [4, 28]. Assumed constant throughout the layers due to strong\nexchange.\ngFrom [46].\nhEstimate obtained from [16].\niEstimate obtained from [16].\njEstimate obtained from [47, 48].\nkSpin current interaction with the 4f moments in Gd is very small [49, 50].\nlChosen such that the simulations \ft the _ qvs.Jexperimental data.\nmNo DMI has been reported at the Co/Gd interface.\na) b) c)\nFig. G5 Comparison of using dynamic and static temperature in the 1D model of CIDWM. The constant temperature is\nchosen to be 85% of the maximum temperature from the COMSOL simulations at a given peak current density, as it leads\nto a comparable shift of the velocity peak. a): 1D CIDWM simulations comparing the velocity for the two temperature\npro\fles at 1.6, 2.6 and 3.6 TA/m2. b): Normalized DW velocity for the 1D CIDWM simulation with the constant and\ndynamic temperature pro\fles, compared with the normalized velocities found from the experiments at peak current density\nJ0= 2:6 TA/m2. c): Same as b), but for J0= 3:6 TA/m2.\nthe current density. As the degree of compensa-\ntion changes, the resulting translation of the DW\nbecomes more of an average over many di\u000berent\nmagnetic compositions which have varying degrees\nof DW mobility.In the experiments a similar phenomenon can\nbe observed, and is qualitatively shown in Fig.\nG5.b and G5.c. Again, for the lower current den-\nsity (J0= 2:6 TA/m2, \fg. G5.b), the heating\ne\u000bect within the temporal width of the current\npulse is limited, and the velocity with thickness is16 Article Title\nwell described by both the constant and dynamic\ntemperature. A drastic change is observed how-\never for large current densities ( J0= 3:6 TA/m2,\n\fg. G5.c), where we \fnd that the dynamic tem-\nperature pro\fle describes the trend of the exper-\niment well. Contrarily the constant temperature\napproximation does not work anymore beyond\npotentially pinpointing the shift in the velocity\npeak.\nReferences\n[1] S. K. Kim, G. S. D. Beach, K.-J. Lee, T. Ono,\nT. Rasing, and H. Yang, \\Ferrimagnetic spin-\ntronics,\" Nature Materials , vol. 21, no. 1,\npp. 24{34, 2022.\n[2] B. Dieny, I. L. Prejbeanu, K. Garello,\nP. Gambardella, P. Freitas, R. Lehndor\u000b,\nW. Raberg, U. Ebels, S. O. Demokritov,\nJ. Akerman, A. Deac, P. Pirro, C. Adel-\nmann, A. Anane, A. V. Chumak, A. Hirohata,\nS. Mangin, S. O. Valenzuela, M. C. Onba\u0018 sl\u0010,\nM. d'Aquino, G. Prenat, G. Finocchio,\nL. Lopez-Diaz, R. Chantrell, O. Chubykalo-\nFesenko, and P. Bortolotti, \\Opportuni-\nties and challenges for spintronics in the\nmicroelectronics industry,\" Nature Electron-\nics, vol. 3, no. 8, pp. 446{459, 2020.\n[3] M. L. Alexey V.Kimel, \\Writing magnetic\nmemory with ultrashort light pulses,\" Nature\nReviews Materials , vol. 4, pp. 189{200, Febru-\nary 2019.\n[4] L. Caretta, M. Mann, F. B uttner, K. Ueda,\nB. Pfau, C. M. G unther, P. Hessing,\nA. Churikova, C. Klose, M. Schneider,\nD. Engel, C. Marcus, D. Bono, K. Bagschik,\nS. Eisebitt, and G. S. D. Beach, \\Fast\ncurrent-driven domain walls and small\nskyrmions in a compensated ferrimagnet,\"\nNature Nanotechnology , vol. 13, no. 12,\npp. 1154{1160, 2018.\n[5] T. A. Ostler, J. Barker, R. F. L. Evans,\nR. W. Chantrell, U. Atxitia, O. Chubykalo-\nFesenko, S. El Moussaoui, L. Le Guyader,\nE. Mengotti, L. J. Heyderman, F. Nolting,\nA. Tsukamoto, A. Itoh, D. Afanasiev, B. A.\nIvanov, A. M. Kalashnikova, K. Vahaplar,J. Mentink, A. Kirilyuk, T. Rasing, and A. V.\nKimel, \\Ultrafast heating as a su\u000ecient stim-\nulus for magnetization reversal in a ferrimag-\nnet,\" Nature Communications , vol. 3, no. 1,\np. 666, 2012.\n[6] M. Beens, M. L. M. Lalieu, A. J. M. Dee-\nnen, R. A. Duine, and B. Koopmans, \\Com-\nparing all-optical switching in synthetic-\nferrimagnetic multilayers and alloys,\" Phys-\nical Review B , vol. 100, pp. 220409{, 12\n2019.\n[7] S. S. P. Parkin, M. Hayashi, and L. Thomas,\n\\Magnetic domain-wall racetrack memory,\"\nScience , vol. 320, p. 190, 04 2008.\n[8] D. Kumar, T. Jin, R. Sbiaa, M. Kl aui,\nS. Bedanta, S. Fukami, D. Ravelosona, S.-\nH. Yang, X. Liu, and S. Piramanayagam,\n\\Domain wall memory: Physics, materials,\nand devices,\" Physics Reports , vol. 958,\npp. 1{35, 2022.\n[9] R. Bl asing, A. A. Khan, P. C. Filippou,\nC. Garg, F. Hameed, J. Castrillon, and\nS. S. P. Parkin, \\Magnetic racetrack mem-\nory: From physics to the cusp of applications\nwithin a decade,\" Proceedings of the IEEE ,\npp. 1{19, 2020.\n[10] I. M. Miron, T. Moore, H. Szambolics,\nL. D. Buda-Prejbeanu, S. Au\u000bret, B. Rod-\nmacq, S. Pizzini, J. Vogel, M. Bon\fm,\nA. Schuhl, and G. Gaudin, \\Fast current-\ninduced domain-wall motion controlled by\nthe Rashba e\u000bect,\" Nature Materials , vol. 10,\nno. 6, pp. 419{423, 2011.\n[11] K.-S. Ryu, L. Thomas, S.-H. Yang, and\nS. Parkin, \\Chiral spin torque at mag-\nnetic domain walls,\" Nature Nanotechnology ,\nvol. 8, no. 7, pp. 527{533, 2013.\n[12] P. P. J. Haazen, E. Mur\u0012 e, J. H. Franken,\nR. Lavrijsen, H. J. M. Swagten, and B. Koop-\nmans, \\Domain wall depinning governed\nby the spin hall e\u000bect,\" Nature Materials ,\nvol. 12, no. 4, pp. 299{303, 2013.\n[13] S.-H. Yang, K.-S. Ryu, and S. Parkin,\n\\Domain-wall velocities of up to 750 m s-1Article Title 17\ndriven by exchange-coupling torque in syn-\nthetic antiferromagnets,\" Nature Nanotech-\nnology , vol. 10, no. 3, pp. 221{226, 2015.\n[14] S. S. P. Parkin, \\Systematic variation of\nthe strength and oscillation period of indi-\nrect magnetic exchange coupling through the\n3d, 4d, and 5d transition metals,\" Physical\nReview Letters , vol. 67, pp. 3598{3601, 12\n1991.\n[15] J. Yu, D. Bang, R. Mishra, R. Ramaswamy,\nJ. H. Oh, H.-J. Park, Y. Jeong, P. Van Thach,\nD.-K. Lee, G. Go, S.-W. Lee, Y. Wang,\nS. Shi, X. Qiu, H. Awano, K.-J. Lee, and\nH. Yang, \\Long spin coherence length and\nbulk-like spin{orbit torque in ferrimagnetic\nmultilayers,\" Nature Materials , vol. 18, no. 1,\npp. 29{34, 2019.\n[16] T. Okuno, D.-H. Kim, S.-H. Oh, S. K.\nKim, Y. Hirata, T. Nishimura, W. S. Ham,\nY. Futakawa, H. Yoshikawa, A. Tsukamoto,\nY. Tserkovnyak, Y. Shiota, T. Moriyama, K.-\nJ. Kim, K.-J. Lee, and T. Ono, \\Spin-transfer\ntorques for domain wall motion in antiferro-\nmagnetically coupled ferrimagnets,\" Nature\nElectronics , vol. 2, no. 9, pp. 389{393, 2019.\n[17] I. Radu, K. Vahaplar, C. Stamm, T. Kachel,\nN. Pontius, H. A. D urr, T. A. Ostler,\nJ. Barker, R. F. L. Evans, R. W. Chantrell,\nA. Tsukamoto, A. Itoh, A. Kirilyuk,\nT. Rasing, and A. V. Kimel, \\Tran-\nsient ferromagnetic-like state mediating\nultrafast reversal of antiferromagnetically\ncoupled spins,\" Nature , vol. 472, no. 7342,\npp. 205{208, 2011.\n[18] A. R. Khorsand, M. Savoini, A. Kirilyuk,\nA. V. Kimel, A. Tsukamoto, A. Itoh, and\nT. Rasing, \\Role of magnetic circular dichro-\nism in all-optical magnetic recording,\" Phys-\nical Review Letters , vol. 108, pp. 127205{, 03\n2012.\n[19] P. Li, M. J. G. Peeters, Y. L. W. van Hees,\nR. Lavrijsen, and B. Koopmans, \\Ultra-low\nenergy threshold engineering for all-optical\nswitching of magnetization in dielectric-\ncoated Co/Gd based synthetic-ferrimagnet,\"\nApplied Physics Letters , vol. 119, p. 252402,2021/12/20 2021.\n[20] J. H. Mentink, J. Hellsvik, D. V. Afanasiev,\nB. A. Ivanov, A. Kirilyuk, A. V. Kimel,\nO. Eriksson, M. I. Katsnelson, and T. Rasing,\n\\Ultrafast spin dynamics in multisublattice\nmagnets,\" Physical Review Letters , vol. 108,\npp. 057202{, 01 2012.\n[21] H. Becker, C. J. Kr uckel, D. V. Thourhout,\nand M. J. R. Heck, \\Out-of-plane focusing\ngrating couplers for silicon photonics integra-\ntion with optical MRAM technology,\" IEEE\nJournal of Selected Topics in Quantum Elec-\ntronics , vol. 26, no. 2, pp. 1{8, 2020.\n[22] E. K. Sobolewska, J. Pelloux-Prayer,\nH. Becker, G. Li, C. S. Davies, C. J. Kr uckel,\nL. A. F\u0013 elix, A. Olivier, R. C. Sousa, I. L.\nPrejbeanu, A. I. Kiriliouk, D. V. Thourhout,\nT. Rasing, F. Moradi, and M. J. R. Heck,\n\\Integration platform for optical switch-\ning of magnetic elements,\" in Proc.SPIE ,\nvol. 11461, 8 2020.\n[23] M. L. M. Lalieu, M. J. G. Peeters, S. R. R.\nHaenen, R. Lavrijsen, and B. Koopmans,\n\\Deterministic all-optical switching of syn-\nthetic ferrimagnets using single femtosecond\nlaser pulses,\" Physical Review B , vol. 96,\npp. 220411{, 12 2017.\n[24] L. Wang, H. Cheng, P. Li, Y. Liu, Y. L. W. v.\nHees, R. Lavrijsen, X. Lin, K. Cao, B. Koop-\nmans, and W. Zhao, \\Picosecond switching\nof optomagnetic tunnel junctions,\" arXiv , 11\n2020.\n[25] M. Beens, M. L. M. Lalieu, R. A. Duine, and\nB. Koopmans, \\The role of intermixing in all-\noptical switching of synthetic-ferrimagnetic\nmultilayers,\" AIP Advances , vol. 9, p. 125133,\n2020/05/14 2019.\n[26] L. Wang, Y. L. W. van Hees, R. Lavri-\njsen, W. Zhao, and B. Koopmans, \\Enhanced\nall-optical switching and domain wall veloc-\nity in annealed synthetic-ferrimagnetic mul-\ntilayers,\" Applied Physics Letters , vol. 117,\np. 022408, 2020/08/12 2020.18 Article Title\n[27] M. L. M. Lalieu, R. Lavrijsen, and B. Koop-\nmans, \\Integrating all-optical switching\nwith spintronics,\" Nature Communications ,\nvol. 10, no. 1, p. 110, 2019.\n[28] R. Bl asing, T. Ma, S.-H. Yang, C. Garg, F. K.\nDejene, A. T. N'Diaye, G. Chen, K. Liu, and\nS. S. P. Parkin, \\Exchange coupling torque in\nferrimagnetic Co/Gd bilayer maximized near\nangular momentum compensation tempera-\nture,\" Nature Communications , vol. 9, no. 1,\np. 4984, 2018.\n[29] E. A. Nesbitt, H. J. Williams, J. H. Wernick,\nand R. C. Sherwood, \\Magnetic moments\nof intermetallic compounds of transition\nand rare-earth elements,\" Journal of Applied\nPhysics , vol. 33, pp. 1674{1678, 2021/12/14\n1962.\n[30] K.-J. Kim, S. K. Kim, Y. Hirata, S.-H.\nOh, T. Tono, D.-H. Kim, T. Okuno, W. S.\nHam, S. Kim, G. Go, Y. Tserkovnyak,\nA. Tsukamoto, T. Moriyama, K.-J. Lee, and\nT. Ono, \\Fast domain wall motion in the\nvicinity of the angular momentum compen-\nsation temperature of ferrimagnets,\" Nature\nMaterials , vol. 16, no. 12, pp. 1187{1192,\n2017.\n[31] L. Avil\u0013 es-F\u0013 elix, L. \u0013Alvaro-G\u0013 omez, G. Li, C. S.\nDavies, A. Olivier, M. Rubio-Roy, S. Auf-\nfret, A. Kirilyuk, A. V. Kimel, T. Rasing,\nL. D. Buda-Prejbeanu, R. C. Sousa, B. Dieny,\nand I. L. Prejbeanu, \\Integration of Tb/Co\nmultilayers within optically switchable per-\npendicular magnetic tunnel junctions,\" AIP\nAdvances , vol. 9, p. 125328, 2020/06/18 2019.\n[32] Y. Yang, R. B. Wilson, J. Gorchon, C.-\nH. Lambert, S. Salahuddin, and J. Bokor,\n\\Ultrafast magnetization reversal by picosec-\nond electrical pulses,\" Science Advances ,\nvol. 3, p. e1603117, 11 2017.\n[33] I. M. Miron, K. Garello, G. Gaudin, P.-\nJ. Zermatten, M. V. Costache, S. Au\u000bret,\nS. Bandiera, B. Rodmacq, A. Schuhl, and\nP. Gambardella, \\Perpendicular switching of\na single ferromagnetic layer induced by in-\nplane current injection,\" Nature , vol. 476,\nno. 7359, pp. 189{193, 2011.[34] A. V. Svalov, A. Fernandez, V. O.\nVas'kovskiy, M. Tejedor, J. M. Baran-\ndiar\u0013 an, I. Orue, and G. V. Kurlyandskaya,\n\\Ferrimagnetic properties of Co/(Gd{\nCo) multilayers,\" Journal of Magnetism\nand Magnetic Materials , vol. 304, no. 2,\npp. e703{e705, 2006.\n[35] K. Cai, Z. Zhu, J. M. Lee, R. Mishra, L. Ren,\nS. D. Pollard, P. He, G. Liang, K. L. Teo, and\nH. Yang, \\Ultrafast and energy-e\u000ecient spin{\norbit torque switching in compensated ferri-\nmagnets,\" Nature Electronics , vol. 3, no. 1,\npp. 37{42, 2020.\n[36] E. Martinez, S. Emori, N. Perez, L. Tor-\nres, and G. S. D. Beach, \\Current-driven\ndynamics of Dzyaloshinskii domain walls in\nthe presence of in-plane \felds: Full micromag-\nnetic and one-dimensional analysis,\" Jour-\nnal of Applied Physics , vol. 115, p. 213909,\n2020/06/06 2014.\n[37] S. A. Siddiqui, J. Han, J. T. Finley, C. A.\nRoss, and L. Liu, \\Current-induced domain\nwall motion in a compensated ferrimagnet,\"\nPhysical Review Letters , vol. 121, pp. 057701{\n, 07 2018.\n[38] Y. Hirata, D.-H. Kim, T. Okuno,\nT. Nishimura, D.-Y. Kim, Y. Futakawa,\nH. Yoshikawa, A. Tsukamoto, K.-J. Kim, S.-\nB. Choe, and T. Ono, \\Correlation between\ncompensation temperatures of magnetization\nand angular momentum in GdFeCo ferri-\nmagnets,\" Phys. Rev. B , vol. 97, p. 220403,\nJun 2018.\n[39] A. Thiaville, S. Rohart, \u0013E. Ju\u0013 e, V. Cros, and\nA. Fert, \\Dynamics of Dzyaloshinskii domain\nwalls in ultrathin magnetic \flms,\" vol. 100,\nno. 5, p. 57002, 2012.\n[40] Y. Guan, Increased e\u000eciency of current-\ninduced chiral domain wall motion by inter-\nface engineering . PhD thesis, Martin-Luther-\nUniversit at Halle-Wittenberg, 2021.\n[41] R. Bl asing, Highly e\u000ecient domain wall\nmotion in ferrimagnetic Bi-layer systems\nat the angular momentum compensationArticle Title 19\ntemperature . PhD thesis, Martin-Luther-\nUniversit at Halle-Wittenberg, 2019.\n[42] C. Kittel, \\On the gyromagnetic ratio\nand spectroscopic splitting factor of ferro-\nmagnetic substances,\" Phys. Rev. , vol. 76,\npp. 743{748, Sep 1949.\n[43] J. Pelzl, R. Meckenstock, D. Spoddig,\nF. Schreiber, J. P\raum, and Z. Frait, \\Spin\norbit-coupling e\u000bects on g-value and damping\nfactor of the ferromagnetic resonance in Co\nand Fe \flms,\" Journal of Physics: Condensed\nMatter , vol. 15, pp. S451{S463, feb 2003.\n[44] W. Low and D. Shaltiel, \\Paramagnetic-\nresonance spectrum of gadolinium in sin-\ngle crystals of thorium oxide,\" Journal of\nPhysics and Chemistry of Solids , vol. 6, no. 4,\npp. 315{323, 1958.\n[45] J. Seib and M. F ahnle, \\Calculation of the\nGilbert damping matrix at low scattering\nrates in Gd,\" Phys. Rev. B , vol. 82, p. 064401,\nAug 2010.\n[46] O. Lutes, J. Holmen, R. Kooyer, and O. Aad-\nland, \\Inverted and biased loops in amor-\nphous Gd-Co \flms,\" IEEE Transactions on\nMagnetics , vol. 13, no. 5, pp. 1615{1617,\n1977.\n[47] J. W. Lee, Y.-W. Oh, S.-Y. Park, A. I.\nFigueroa, G. van der Laan, G. Go, K.-J. Lee,\nand B.-G. Park, \\Enhanced spin-orbit torque\nby engineering Pt resistivity in Pt =Co=AlOx\nstructures,\" Phys. Rev. B , vol. 96, p. 064405,\nAug 2017.\n[48] K. Ando, S. Takahashi, K. Harii, K. Sasage,\nJ. Ieda, S. Maekawa, and E. Saitoh, \\Elec-\ntric manipulation of spin relaxation using the\nspin hall e\u000bect,\" Phys. Rev. Lett. , vol. 101,\np. 036601, Jul 2008.\n[49] X. Jiang, L. Gao, J. Z. Sun, and S. S. P.\nParkin, \\Temperature dependence of current-\ninduced magnetization switching in spin\nvalves with a ferrimagnetic CoGd free layer,\"\nPhys. Rev. Lett. , vol. 97, p. 217202, Nov 2006.[50] B. Dieny, \\Giant magnetoresistance in spin-\nvalve multilayers,\" Journal of Magnetism and\nMagnetic Materials , vol. 136, no. 3, pp. 335{\n359, 1994." }, { "title": "1503.00589v2.Frustrated_mixed_spin_1_2_and_spin_1_Ising_ferrimagnets_on_a_triangular_lattice.pdf", "content": "arXiv:1503.00589v2 [cond-mat.stat-mech] 1 Jun 2015Frustrated mixed spin-1/2 and spin-1 Ising ferrimagnets on a\ntriangular lattice\nM.ˇZukoviˇ c∗and A. Bob´ ak\nDepartment of Theoretical Physics and Astrophysics, Facul ty of Science,\nP. J.ˇSaf´ arik University, Park Angelinum 9, 041 54 Koˇ sice, Slov akia\n(Dated: July 16, 2018)\nAbstract\nMixed spin-1/2 and spin-1 Ising ferrimagnets on a triangula r lattice with sublattices A, B and\nC are studied for two spin value distributions ( SA,SB,SC) = (1/2,1/2,1) and (1 /2,1,1) by Monte\nCarlo simulations. The non-bipartite character of the latt ice induces geometrical frustration in\nboth systems, which leads to the critical behavior rather di fferent from their ferromagnetic coun-\nterparts. We confirm second-order phase transitions belong ing to the standard Ising universality\nclass occurring at higher temperatures, however, in both mo dels these change at tricritical points\n(TCP) to first-order transitions at lower temperatures. In t he model (1 /2,1/2,1), TCP occurs on\nthe boundary between paramagnetic and ferrimagnetic ( ±1/2,±1/2,∓1) phases. The boundary\nbetween two ferrimagnetic phases ( ±1/2,±1/2,∓1) and (±1/2,∓1/2,0) at lower temperatures is\nalways first order and it is joined by a line of second-order ph ase transitions between the paramag-\nnetic and the ferrimagnetic ( ±1/2,∓1/2,0) phases at a critical endpoint. The tricritical behavior\nis also confirmed in the model (1 /2,1,1) on the boundary between the paramagnetic and ferrimag-\nnetic (0,±1,∓1) phases.\nPACS numbers: 05.50.+q, 64.60.De, 75.10.Hk, 75.30.Kz, 75.50.Gg\nKeywords: Mixed-spin system, Frustrated Ising ferrimagnet, Tr iangular lattice, Monte Carlo simulation,\nTricritical point, Critical endpoint\n1I. INTRODUCTION\nMixed-spin Ising systems have been mostly investigated as possible m odels of some\ntypesofferrimagneticandmolecular-basedmagneticmaterials. Th eusedapproachesinclude\nanexacttreatmentinspecialcases1–5,mean-fieldapproximation6,7,effective-fieldtheorywith\ncorrelations8–14, Monte Carlo simulations15–22and some other methods23–28. The main focus\nwere their phase diagrams as well as technologically interesting comp ensation behavior with\npossibility to achieve zero total magnetization by tuning of tempera ture below the critical\npoint. Most of the studies considered the simplest models consisting of two sublattices one\nof which is occupied with spins S= 1/2 and the other with S= 1. Such a mixed-spin model\ncan be described by the Hamiltonian\nH=−J/summationdisplay\n/angbracketlefti,j/angbracketrightσiSj−D/summationdisplay\njS2\nj, (1)\nwhereσi=±1/2 andSj=±1,0 are spins on different sublattices, ∝angbracketlefti,j∝angbracketrightdenotes the sum\nover nearest neighbors, J <0 is a antiferromagnetic exchange interaction parameter and D\nis a single-ion anisotropy parameter. Negative values of the parame terDfavor nonmagnetic\nstates with Sj= 0 and positive values magnetic states with Sj=±1.\nDue to persisting ambiguities majority of the investigations focused on the simplest lat-\ntices, i.e., the square in two and cubic in three dimensions. We note tha t a long standing\ncontroversy regarding the critical and compensation behaviors e ven for the most studied\ncase of the model on a square lattice was solved only recently by Mon te Carlo simulation\nthat has convincingly shown22that there are neither tricritical nor compensation points, as\nhad been suggested by some previous approximative approaches6,8,9,28. On the other hand,\nin the same study the presence of both the tricritical point and a line of compensation\npoints was confirmed in the three-dimensional model on a simple cubic lattice. This finding\nmight suggest that the increased dimensionality is responsible for th e appearance of the tri-\ncritical and compensation behaviors. Nevertheless, our recent s tudy on a triangular lattice\nferromagnet29demonstrated that the tricritical point can also appear in a two-dim ensional\nlattice as long as the coordination number is sufficiently high. The effec t of the coordination\nnumber in the presence of bond disorder on tricritical behavior of a two dimensional system\nwas also recently studied in a random Blume-Capel model on a triangu lar lattice30.\nWe point out that the previous studies were performed on bipartite lattices, in which\n2A BC(a)\nAC(b)\nB\nFIG. 1: (Color online) Mixed-spin S= (SA,SB,SC) models on a triangular lattice consisting of\nsublattices A, B and C, with (a) S= (1/2,1/2,1) mixing - model I and (b) S= (1/2,1,1) mixing\n- model II. Small and large circles denote spin-1/2 and spin- 1 sites, respectively.\ncase the sign of the exchange interaction is irrelevant to the therm odynamic and critical\nproperties of the model in the absence of an external field. On the other hand, the present\nmixed-spin model is considered on a non-bipartite triangular lattice, in which case the sign\nof the exchange interaction matters. Namely, in contrast to the f erromagnetic case, the\nferrimagnetic interaction will induce geometrical frustration, whic h can be expected to have\nsome impact on the critical behavior. As shown in Fig. 1, the lattice co nsists of three\nsublattices A, B and C, occupied with spins S= (SA,SB,SC). This allows to further study\nthe model in two different mixing modes. We can consider a mixed-spin S= (1/2,1/2,1)\nmodel I, as schematically depicted in Fig. 1(a), in which one sublattice is occupied with\nspinS= 1 sites and the remaining two sublattices with spin S= 1/2 sites. Thus, each\nspin-1 site is surrounded by z= 6 nearest neighbors with spin S= 1/2. The other way\nof the spin-mixing is realized in a S= (1/2,1,1) model II, shown in Fig. 1(b), which is\nobtained when the spin-1/2 and spin-1 sites in the model I are swapp ed. The two models\nwere shown to display qualitatively different critical behaviors even f or the ferromagnetic\nexchange interactions29.\nThe goal of the present study is to examine effects of the geometr ical frustration on the\ncritical behavior of the above defined ferrimagnetic mixed-spin sys tems, to determine their\nphase diagrams and to confront them with their ferromagnetic cou nterparts as well as the\npure spin-1/2 and spin-1 antiferromagnetic systems.\n3100102104106−1−0.8−0.6−0.4−0.200.20.40.60.81\nMCSe/|J| mO\n \nmodel II: ms3, kBT/|J|=1.0model I: ms1, kBT/|J|=0.7\nmodel I: e/|J|, kBT/|J|=0.7\nmodel II: e/|J|, kBT/|J|=1.0\nFIG. 2: (Color online) Time evolutions of the internal energ y per site e/|J|and the staggered mag-\nnetizations ms1,ms3(see definitions below), starting from thedisordered phase at the temperatures\nkBT/|J|= 0.7 in model I and 1.0 in model II, for D/|J|= 0 and L= 120.\nII. MONTE CARLO SIMULATION\nIn order to study the behavior of various thermodynamic quantitie s in the parameter\nspace and to determine the phase diagrams we use Monte Carlo (MC) simulations with the\nMetropolis update rule and employ the periodic boundary conditions. We consider lattices\nwith the size L×L, withLranging from 24 up to 120. We perform N= 2×105up\nto 106MCS (Monte Carlo sweeps), the first 20% of which are used to bring t he system\nto equilibrium and then discarded, and the remaining data are used to estimate thermal\naverages and statistical errors. In order to demonstrate that the used MCS is sufficient to\nensure equilibrium conditions, in Fig. 2 we present MC time evolutions of some relevant\nquantities, such as the order parameters ms1(model I) and ms3(model II) and the internal\nenergy per site e/|J|, starting from the disordered phase. The simulations are perform ed\nin both models for the largest of the considered lattice sizes L= 120, which is the most\ndifficult to equilibrate, at the temperatures in the vicinity of the resp ective critical points\nforD/|J|= 0. The plots show that in these cases the equilibrium is reached in less than\n104MCS.\nThe phase boundaries are roughly determined from the maxima of so me thermodynamic\nfunctions, such as the specific heat, for a selected fixed value of L. We chose L= 48, as\n4a compromise value above which the specific heat maxima positions do n ot change con-\nsiderably and the phase diagrams can be determined in a relatively wide parameter space\nin a reasonable computational time. In the region where the critical line as a function of\nthe single-ion anisotropy parameter Dis more or less horizontal it is convenient to obtain\ntemperature dependencies of the calculated quantities at a fixed v alue ofD. In such a\ncase simulations start from the paramagnetic phase using random in itial configurations with\nthe temperature gradually decreased and a new simulation starting from the final config-\nuration obtained at the previous temperature. On the other hand , if the phase boundary\nshape changes to vertical we obtain variations of the quantities as functions of the single-ion\nanisotropy parameter Dat a fixed temperature. Then simulations start from appropriately\nchosen states (i.e., not necessarily random), expected in the cons idered region of the param-\neter space. Following the above described approach we ensure tha t the system is maintained\nclose to the equilibrium in the entire range of the changing parameter and thus considerably\nshortens thermalization periods. In order to estimate statistical errors, we perform three\nindependent simulations at all considered parameter values.\nAt some selected points of the phase boundaries we perform a more thorough finite-\nsize scaling (FSS) analysis in order to determine more precisely the loc ation of the critical\npoints and the corresponding critical exponents. In such a case w e perform more extensive\nsimulations using up to N= 107MCS and apply the reweighing techniques31. For more\nreliableestimationofstatistical errors, inthiscase weusedtheΓ-m ethod32. Having obtained\nthe maxima of the relevant quantities, we apply the linear fitting proc edure for logarithms of\ndata with errors, following the method in York et al.33. In order to assess the quality of the\nfitting, asameasureofgoodnessoffitweevaluatedanadjustedc oefficient ofdetermination34\nof the linear fit R2. The critical points and the exponents are then extracted from t he FSS\nanalysis, using the linear sizes L= 24,48,72,96 and 120.\nWe calculate the following quantities: the internal energy per spin e=∝angbracketleftH∝angbracketright/L2, the\nrespective sublattice magnetizations per site mX, (X = A, B or C), as order parameters on\nthe respective sublattices, which for the model I are given by\nmA(B)= 3∝angbracketleft|MA(B)|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\ni∈A(B)σi/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2, (2)\nmC= 3∝angbracketleft|MC|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\ni∈CSi/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2, (3)\n5and for the model II by\nmA= 3∝angbracketleft|MA|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\ni∈Aσi/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2, (4)\nmB(C)= 3∝angbracketleft|MB(C)|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\ni∈B(C)Si/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2, (5)\nwhere∝angbracketleft···∝angbracketrightdenotes thermal average. Based on the ground-state consider ations (see below),\nfortheidentified orderedphaseswe additionallydefinethefollowingo rder parametersforthe\nentire system, which take values between 0 in the fully disordered an d 1 in the fully ordered\nphase. For the model I we introduce two order parameters (stag gered magnetizations per\nsite)ms1andms2given by\nms1=∝angbracketleft|Ms1|∝angbracketright/L2=/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle2/summationdisplay\ni∈Aσi+2/summationdisplay\nj∈Bσj−/summationdisplay\nk∈CSk/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2, (6)\nand\nms2=∝angbracketleft|Ms2|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\ni∈Aσi−/summationdisplay\nj∈Bσj/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/L2. (7)\nFor the model II we define the order parameter ms3as\nms3=∝angbracketleft|Ms3|∝angbracketright/L2= 3/angbracketleftBig/vextendsingle/vextendsingle/vextendsingle/summationdisplay\nj∈BSj−/summationdisplay\nk∈CSk/vextendsingle/vextendsingle/vextendsingle/angbracketrightBig\n/2L2. (8)\nUnlike in ferrimagnetic systems on bipartite lattices, the three subla ttices in the present\nsystem facilitate spin arrangements in such a way that the total ne t magnetization is always\nzero and, therefore, of no practical use.\nFurther, we calculate the susceptibilities pertaining to the respect ive order parameters\nO=MX(X = A, B, C) and also O=Msi(i= 1,2 or 3)\nχO=∝angbracketleftO2∝angbracketright−∝angbracketleftO∝angbracketright2\nNOkBT, (9)\nthe specific heat per site c\nc=∝angbracketleftH2∝angbracketright−∝angbracketleftH∝angbracketright2\nNOkBT2, (10)\nwhereNOis the number of sites on the (sub)lattice on which Ois defined. Further, we\ndefine the logarithmic derivatives of ∝angbracketleftO∝angbracketrightand∝angbracketleftO2∝angbracketrightwith respect to β= 1/kBT,\nD1O=∂\n∂βln∝angbracketleftO∝angbracketright=∝angbracketleftOH∝angbracketright\n∝angbracketleftO∝angbracketright−∝angbracketleftH∝angbracketright, (11)\n6D2O=∂\n∂βln∝angbracketleftO2∝angbracketright=∝angbracketleftO2H∝angbracketright\n∝angbracketleftO2∝angbracketright−∝angbracketleftH∝angbracketright, (12)\nand finally the fourth-order Binder cumulant UOcorresponding to the quantity O\nUO= 1−∝angbracketleftO4∝angbracketright\n3∝angbracketleftO2∝angbracketright2. (13)\nThe above standard thermodynamic quantities (9-10), as well as t he less traditional ones\ndefined by Eqs. (11-13), serve to obtain estimates of the respec tive critical exponents by FSS\nanalysis35,36. In particular, we use the following scaling relations, applied to the ma ximum\nvalues of the following functions:\nχO,max(L)∝LγO/νO, (14)\ncmax(L)∝Lα/νO, (15)\nD1O,max(L)∝L1/νO, (16)\nD2O,max(L)∝L1/νO, (17)\nwhereαis the critical exponent of the specific heat and νO,γOare the critical exponents\nof the correlation length and susceptibility, respectively, pertainin g to the quantity O. In\ncase of a first-order phase transition, the above quantities (14- 17) are expected to scale as\n∝Ld, whered= 2 is the system dimension. The order parameter cumulant, defined by\nEq. (13), can also serve for a simple yet relatively precise location of the phase transition\npoint as a point at which the cumulant curves obtained for different s ystem sizes intersect\nat a universal value, e.g., UO(Tc) = 0.611 for a two-dimensional Ising model37.\nIII. RESULTS\nA. Ground state\nLet us first identify all the possible ground states (GS) for entire r ange of the single-ion\nanisotropy parameter D. Considering the lattice system consisting of three interpenetrat ing\nsublatticesA,BandC,asschematicallydepictedinFig.1, theHamilton iansoftherespective\nmodels I and II can be defined as\nHI=−J/parenleftBig/summationdisplay\ni∈A,j∈Bσiσj+/summationdisplay\ni∈A,k∈CσiSk+/summationdisplay\nj∈B,k∈CσjSk/parenrightBig\n−D/summationdisplay\nk∈CS2\nk, (18)\n7TABLE I: Ground state configurations and the respective ener gies for different ranges of the single-\nion anisotropy parameter.\nModel I II\nD/|J| State Energy e/|J|State Energy e/|J|\n(−∞,−3/2) FRI\n2: (±1/2,∓1/2,0) −1/4 P: (0 ,0,0) 0\n(−3/2,∞) FRI\n1: (±1/2,±1/2,∓1)−3/4−D/3|J|FRII: (0,±1,∓1)−1−2D/3|J|\nHII=−J/parenleftBig/summationdisplay\ni∈A,j∈BσiSj+/summationdisplay\ni∈A,k∈CσiSk+/summationdisplay\nj∈B,k∈CSjSk/parenrightBig\n−D/parenleftBig/summationdisplay\nj∈BS2\nj+/summationdisplay\nk∈CS2\nk/parenrightBig\n.(19)\nFocusing on a triangular elementary unit cell consisting of the spins SA,SB,SC, from\nthe Hamiltonians (18) and (19) one can obtain expressions for the r educed energies per\nspine/|J|of different spin arrangements as functions of D/|J|. Then the ground states\nare determined as configurations corresponding to the lowest ene rgies for different values of\nD/|J|, as tabulated in Table I. There are two long-range order (LRO) fer rimagnetic (FR)\nstates FRI\n1and FRI\n2in the model I and one LRO ferrimagnetic state FRIIand one disordered\nparamagnetic (P) phase in the model II. We note that while in FRI\n2zero means nonmagnetic\nstates (Sj= 0) of spins on sublattice C, in FRIIzero means magnetic states ( σi=±1/2) of\nspins on sublattice A but equally in states +1 /2 and−1/2, thus giving zero net sublattice\nmagnetization. The critical value of the single-ion anisotropy param eter separating the\nrespective phases is the same for both models Dc/|J|=−3/2.\nB. Monte Carlo\n1. Model I: S= (1/2,1/2,1)\nPhase boundaries between the disordered paramagnetic and the r espective ordered ferri-\nmagnetic phases are determined from the specific heat maxima for a fixedL= 4848and the\nnature of the ordered phases is established from the introduced o rder parameters ms1and\nms2. TheseareshowninFigs.3(a)and3(b)forselectedvaluesofthep arameter D/|J|below,\nclose to and above the critical value Dc/|J|. All the specific heat curves show pronounced\nsharp peaks, signifying phase transitions to the low-temperature ferrimagnetic states. How-\never, the (pseudo)transition temperatures appear to be a nonm onotonic functions of D/|J|.\n80.20.30.40.50.60.70.80.900.511.522.5\n \nkBT/|J|c(a)\nD/|J|=0 D/|J|=−2D/|J|=−1.48\n0.20.30.40.50.60.70.80.900.10.20.30.40.50.60.70.80.91\nkBT/|J|ms1, ms2\n \n(b)\nms1, D/|J|=0\nms2, D/|J|=0\nms1, D/|J|=−2ms2, D/|J|=−2ms1, D/|J|=−1.48\nms2, D/|J|=−1.48\nFIG. 3: (Color online) Temperature dependencies of (a) the s pecific heat and (b) the order param-\neters, for selected values of D/|J|and a fixed lattice size L= 48. In (a) additionally the results\nforL= 72 are shown (squares).\nMore specifically, the transition temperature for D/|J|=−1.48, i.e., close to the critical\nvalueDc/|J|, is lower than for the other two values below and above Dc/|J|. Moreover, the\ncorresponding specific heat maximum has a spike-like shape and its ma gnitude is one order\nhigher then the other maxima, which is typical for a first-order pha se transition. In order\nto support our claim, that the peaks’ positions do not significantly c hange above L= 48,\nin Fig. 3(a) we also included the results obtained for L= 72. This will also become evi-\ndent later on in the respective phase diagrams in which for some selec ted points the phase\ntransition temperatures estimated for L= 48 will be compared with those determined for L\nextrapolated to infinity. The order parameters depicted in Fig. 3(b ) demonstrate that the\ntransition for D/|J|=−2 is to the state ( ±1/2,∓1/2,0), characterized by a finite values\nofms2, while for D/|J|=−1.48 and 0 the system tends to the state ( ±1/2,±1/2,∓1),\ncharacterized by a finite values of ms1. The discontinuous behavior of the order parameter\nforD/|J|=−1.48 corroborates the first-order nature of the transition.\nBy FSS analysis at two representative values of D/|J|= 0 and −2, selected on either\nside of the critical value Dc/|J|=−3/2, we confirmed that the disorder-to-order phase\ntransitions to both ( ±1/2,±1/2,∓1) and (±1/2,∓1/2,0) ferrimagnetic phases are indeed\nsecond order and belong to the standard Ising universality class. T he slopes of the fitted\n93 3.5 4 4.5 522.533.544.555.566.5\nln(L) \n γM\ns1/νM\ns1 = 1.756 ± 0.006\n 1/νM\ns1 = 0.989 ± 0.011\n 1/νM\ns1 = 0.988 ± 0.013ln(χM\ns1,max), ln(D1M\ns1,max), ln(D2M\ns1,max)ln(χM\ns1,max)\nln(D1M\ns1,max)\nln(D2M\ns1,max)(a)\n3 3.5 4 4.5 51234567\nln(L)ln(χM\ns2,max), ln(D1M\ns2,max), ln(D2M\ns2,max)\n \n γM\ns2/νM\ns2=1.753 ± 0.003\n 1/νM\ns2 = 0.982 ± 0.008 1/νM\ns2 = 0.985 ± 0.008ln(χM\ns2,max)\nln(D1M\ns2,max)\nln(D2M\ns2,max)(b)\n0.745 0.75 0.7550.520.540.560.580.60.620.64\nkBT/|J|UM\ns1\n \nL=24\nL=48\nL=72\nL=96\nL=120(c)\n0.362 0.364 0.366 0.368 0.370.3720.450.50.550.60.65UM\ns2\nkBT/|J| \nL=24\nL=48\nL=72\nL=96\nL=120(d)\n−1−0.5 00.511.522.50.20.250.30.350.40.450.50.550.60.65\nL1/ν\nM\ns1(T−Tc)/TcUM\ns1\n \nL=24\nL=48\nL=72\nL=96\nL=120(e)\n−1−0.5 00.511.522.50.20.250.30.350.40.450.50.550.60.65\nL1/ν\nM\ns2(T−Tc)/TcUM\ns2\n \nL=24\nL=48\nL=72\nL=96\nL=120(f)\nFIG. 4: (Color online) FSS analysis of the critical exponent s ratios 1 /νOandγO/νO(a,b), the\nfourth-order cumulant UOtemperature dependencies for different L(c,d) and the UOdata collapse\nanalysis (e,f), for O=Ms1atD/|J|= 0 (left column) and for O=Ms2atD/|J|=−2 (right\ncolumn). The coefficients of determination R2for the respective fits (from top to bottom) are\n0.9999, 0.9998, 0.9997 in (a) and 0.9999, 1.0000, 1.0000 in ( b).\n10−1.8−1.7−1.6−1.5−1.4−1.3−1.2−1.100.20.40.60.81\nD/|J|ms1\n \nD/|J| down\nD/|J| upkBT/|J|=0.150.20.250.30.35\nFIG. 5: (Color online) Order parameter ms1as a function of the increasing ( ⊲) and decreasing ( ⊳)\nsingle-ion anisotropy parameter D/|J|at various temperatures and L= 48. The double-headed\narrows mark the hysteresis widths.\ncurves in Figs. 4(a) and 4(b) represent ratios of the critical expo nents 1/νOandγO/νO,\nfollowing from the scaling relations (14)-(17), where O=Msi(i= 1,2). We also checked\nthat for both D/|J|= 0 and −2 the specific heat maxima follow the logarithmic scaling\ncmax=c0+c1ln(L), as expected for the Ising universality class in two dimensions (not\nshown). Having performed the FSS analysis, these two critical poin ts can be determined\nwith a higher accuracy from the Binder cumulant crossing method38, as an intersection of\nthe Binder parameter UOcurves for different lattice sizes LandO=Msi(i= 1,2) (see\nFigs.4(c)and4(d)). Thecriticaltemperaturesweredetermined askBTc/|J|= 0.7485±0.001\natD/|J|= 0 and kBTc/|J|= 0.3663±0.001 atD/|J|=−2. Also the critical values of\nthe Binder cumulants UO(Tc) = 0.61137confirm the Ising universality class. Furthermore, in\nFigs. 4(e),4(f)we show thatfor thecritical temperatures deter mined by theBinder cumulant\ncrossing method the data for different lattice sizes indeed collapse o n a single curve. The\napparent first-order character of the phase transition at D/|J|=−1.48 will be discussed\nbelow.\nLet us now examine the transition between the two low-temperatur e ferrimagnetic phases\nFRI\n1and FRI\n2. Since the expected phase boundary is almost vertical to the x-ax is, instead\nof the temperature dependencies of various thermodynamic func tions it is more convenient\nto look into their single-ion parameter dependencies at a fixed tempe rature. By plotting\n11−0.25 −0.2 −0.15 −0.1 −0.0500.511.522.533.544.55x 105\ne/|J| \nL=24\nL=48\nL=96\nL=120(a)\n3 3.5 4 4.5 5123456789\nln(L)ln(cmax) ln( χM\ns1,max)\n \nD/|J|=−1.47\nD/|J|=−1.47D/|J|=−1.48, αt/νt=1.60D/|J|=−1.4825, α/ν=2D/|J|=−1.48, γt/νt=1.85(b)\nD/|J|=−1.4825, γ/ν=2\n3 3.5 4 4.5 51234567\nln(L) \n γM\ns1/νM\ns1 = 1.744 ± 0.008\n 1/νM\ns1 = 0.996 ± 0.009\n 1/νM\ns1 = 0.999 ± 0.007ln(χM\ns1,max), ln(D1M\ns1,max), ln(D2M\ns1,max)ln(χM\ns1,max)\nln(D1M\ns1,max)\nln(D2M\ns1,max)(c)\nFIG. 6: (Color online) (a) Energy distributions for D/|J|=−1.47 and different L. The respective\ntemperatures are tuned by the reweighing technique to achie ve approximately equal peak heights.\n(b) FSS analysis of the susceptibility χMs1and the specific heat c, forD/|J|=−1.47,−1.48 and\n−1.4825. For D/|J|=−1.48and−1.4825thelog-log plotsarerespectivelyfittedtotheexactva lues\nof the tricritical exponents ratios γt/νt= 1.85,αt/νt= 1.60 and the exponent 2, corresponding\nto the system volume. (c) FSS analysis of χMs1,D1Ms1,D2Ms1forD/|J|=−1.46, with the\ncoefficients of determination R2for the respective fits (from top to bottom) 0.9999, 1.0000, 0 .9999.\nthe order parameters as increasing and decreasing functions of D/|J|one can observe their\ndiscontinuous character and the appearance of hysteresis loops , the widths of which increase\nwith decreasing temperature. This behavior is demonstrated in Fig. 5 for the order parame-\nterms1andL= 48. Such behaviorsignalsfirst-orderphasetransitions. Theyse em topersist\n124 4.5 5 5.5 6−0.500.51\n105 MCSsublattice magnetizations\n \nm1\nm2\nm3(a)\n3.5 4 4.5 5 5.5−1−0.8−0.6−0.4−0.200.20.40.60.81\n105 MCSsublattice magnetizations\n \nm1\nm2\nm3(b)\nFIG. 7: (Color online) Time evolution of the sublattice magn etizations at the points (a)\n(D/|J|,kBT/|J|) = (−1.487,0.33) and (b) ( −1.47,0.35), forL= 48.\neven to higher temperatures at which the hysteretic behavior in no t apparent any longer.\nThe highest temperatures at which we still could observe some signs of first-order phase\ntransitions, such as bimodal energy distribution, was kBT/|J|= 0.35 and the corresponding\nvalue ofD/|J|=−1.47 (see Fig. 6(a)).\nNevertheless, the energy barrier separating the two peaks upon initial increase seems\nto decrease for larger L, which indicate that the phase transition may not be truly first\norder. Indeed, when we checked whether the specific heat and su sceptibility scale with the\nsystem volume, as it should be in case of a first-order transition, we found that within\nthe used lattice sizes the linear ansatz could not be established for D/|J|=−1.47, as\nshown in Fig. 6(b). We note that such a behavior that can lead to misin terpretation of a\nsecond-order transition as first order was also observed in the Blu me-Capel model as well\nas in the frustrated J1−J2Ising antiferromagnet on a square lattice39,40. Clearly first-order\nscaling is observed only at D/|J|=−1.4825. Nevertheless, fairly good linear fits were also\nachieved in the case of D/|J|=−1.48 with the exponents that are between 2 and the exact\ntricritical values (1.85 for γ/νand 1.6 for α/ν)41,42, suggesting that the system is close to\nthe tricritical point. On the other hand, for D/|J|=−1.46, the FSS presented in Fig. 6(c)\nclearly indicates a second-order phase transition. Thus, we rough ly estimate the tricritical\npoint at ( Dt/|J|,kBTt/|J|) = (−1.47±0.01,0.35±0.01).\nIt is interesting to notice that the phase transition at this point is no longer between\n13−3 −2 −1 0 1 200.10.20.30.40.50.60.70.80.9\nD/|J|kBTc/|J|P\nCEFR1I\nFR2ITCP\nfirst order\nFIG. 8: (Color online) Phase diagram of the model I in ( kBT/|J|−D/|J|) parameter space. The\nempty circles represent thephasetransition temperatures kBTc/|J|between theparamagnetic state\nP and the ferrimagnetic states FRI\n1(±1/2,±1/2,∓1) and FRI\n2(±1/2,∓1/2,0), estimated from the\nspecific heat peaks for L= 48, the empty triangles mark the hysteresis widths at first- order\ntransitions between the phases FRI\n1and FRI\n2with the expected phase transition boundary marked\nby the dash-dot line. The filled symbols at finite temperature s show more precise values obtained\nfrom the FSS analysis and the Binder cumulant crossing, wher e the diamond is the tricritical point\n(TCP), the hexagon is the critical endpoint (CE) and the squa re at (Dc/|J|,kBTc/|J|) = (−3/2,0)\nrepresents the exact value of the GS transition point.\nthe two ferrimagnetic phases FRI\n1and FRI\n2but between the phase FRI\n1and the param-\nagnetic phase. This is in line with the above observation of the first-o rder-like features\nof the phase transition at D/|J|=−1.4825 between the paramagnetic and FRI\n1phases.\nThe fact that there are first-order phase transitions between F RI\n1and FRI\n2as well as FRI\n1\nand paramagnetic phases can be verified by looking at the relevant o rder parameters. For\ndemonstration, in Figs. 7 we show segments of sublattice magnetiza tion time series ob-\ntained at ( D/|J|,kBT/|J|) = (−1.487,0.33) (Fig. 7(a)) and ( −1.47,0.35) (Fig. 7(b)). In\nboth cases we can see discontinuous switching between two phases , however, in either\npoint those phases are different. Namely, at ( −1.487,0.33) the sublattice magnetizations\n(mA,mB,mC) switch between the values characteristic for the states ( ±1/2,∓1/2,0)\nand (±1/2,±1/2,∓1), while at ( −1.47,0.35) they switch between the values characteris-\n14tic for the states (0 ,0,0) and (±1/2,±1/2,∓1). This finding implies the existence of a\ncritical endpoint (CE) at which the second-order transition bound ary between the para-\nmagnetic and FRI\n2phases joins the above discussed first-order transition line at abo ut\n(Dce/|J|,kBTce/|J|) = (−1.485±0.005,0.335±0.005).\nThe resulting phase diagram is presented in Fig. 8. The empty circles r epresent the\npseudo-critical points determined from the specific heat maxima at L= 48 and the empty\ntriangles mark the metastable branches of the first-order phase transitions obtained from\nthe order parameter hysteresis loops. The filled circles and square s at finite temperatures\nrepresent respectively second- and first-order transition point s obtained from the FSS anal-\nysis and the filled diamond marks approximate location of the tricritica l point. The filled\nsquare at ( Dc/|J|,kBTc/|J|) = (−3/2,0)shows the exact locations ofthe ground-statephase\ntransition. The expected first-order phase transition boundary , marked by the dash-dot line,\nis obtained by a simple linear interpolation between the estimated tricr itical and the exact\nGS transition points and only serves as a guide to the eye. We note th at in this highly sup-\npressed mixed-phase region the first-order phase boundaries ca n be located quite precisely,\nfor example, by multicanonical MC simulations43,44. However, in our case, we are only in-\nterested in approximate location of the phase boundaries. Then, c onsidering the exact value\nof the GS transition point, the estimate location of the tricritical po int and assuming no\nanomalous behavior, such as reentrance, the low-temperature p art of the phase boundary\nmust be practically vertical to the D/|J|axis.\n2. Model II: S= (1/2,1,1)\nAlso for the model II the phase boundary as a function of the single -ion anisotropy\nparameter D/|J|is estimated from the specific heat maxima for L= 48, except for D/|J|\nclose to the critical value of Dc/|J|=−3/2, where the boundary becomes almost vertical.\nThe only identified LRO phase is the ferrimagnetic phase FRII: (±1/2,±1,∓1), present\nforD > D c, and the phase transition is second order complying with the standa rd Ising\nuniversality class. This is illustrated in Fig. 9(a) for a selected value of D/|J|= 0, in which\nwe show that the obtained ratios of the critical exponents 1 /νMs3andγMs3/νMs3are in a\ngood agreement with the 2D Ising universality class values 1 and 7 /4, respectively. We also\nchecked the consistency of the specific heat critical exponent va lueα= 0 by verifying the\n153 3.5 4 4.5 522.533.544.555.5\nln(L)ln(χM\ns3,max), ln(D1M\ns3,max), ln(D2M\ns3,max)\n \nln(χM\ns3,max)\nln(D1M\ns3,max)\nln(D2M\ns3,max)(a)\nγM\ns3/νM\ns3 = 1.757 ± 0.006\n1/νM\ns3 = 1.000 ± 0.013\n1/νM\ns3 = 1.001 ± 0.015\n1.055 1.061.065 1.071.075 1.080.580.60.620.640.660.68\nkBT/|J|UM\ns3\n \n0 0.02 0.041.061.081.11.121.141.161.18\nL−1 \nkBTmax,χ\nM\ns3/|J|\nkBTmax,D1\nM\ns3/|J|\nkBTmax,D2\nM\ns3/|J|\nL=24\nL=48\nL=72\nL=96\nL=120(b)\n−1−0.500.511.520.20.250.30.350.40.450.50.550.60.65\nL1/ν\nM\ns3(T−Tc)/TcUM\ns3\n \nL=24\nL=48\nL=72\nL−96\nL=120(c)\nFIG. 9: (Color online) (a) FSS analysis of the critical expon ent ratios 1 /νMs3andγMs3/νMs3, (b)\nthe fourth-order cumulant UMs3temperature dependencies for different L, and (c) the UMs3data\ncollapse analysis at D/|J|= 0. In (a) the coefficients of determination R2for the respective fits\n(from top to bottom) are 0.9998, 0.9999 and 0.9999. Inset in ( b) shows an alternative way of the\ncritical temperature estimation from the FSS analysis.\nlogarithmic scaling (not shown). The critical temperature for D/|J|= 0 estimated by the\nBinder cumulant method (Fig. 9(b)) and FSS analysis (inset in Fig. 9(b )) takes the value\nkBTc/|J|= 1.0635±0.0015 and the cumulant curves for different Lintersect at the universal\nvalue ofUMs3(Tc) = 0.611.\nOn approach to the critical value Dc/|J|=−3/2 the phase boundary rapidly drops and\n16−1.6 −1.55 −1.5 −1.45 −1.4 −1.3500.10.20.30.40.50.60.70.80.91\nD/|J|ms3kBT/|J|=0.3\nkBT/|J|=0.05kBT/|J|=0.2\nFIG. 10: (Color online) Order parameter ms3as a function of the increasing ( ⊲) and decreasing\n(⊳) single-ion anisotropy parameter D/|J|at various temperatures and L= 48.\nbecomes almost vertical. Therefore, in order to locate the critical temperatures in this re-\ngion, it is more convenient to measure the physical quantities at a fix ed temperature as\nfunctions of the parameter D/|J|. At sufficiently low temperatures the measured quantities\nshow some properties typical for first-order phase transitions. Namely, as the anisotropy\nparameter D/|J|is decreased and increased at the fixed temperature the sublattic e mag-\nnetizations, the order parameter ms3and the internal energy show discontinuities at some\nvalues of D/|J|, as demonstrated in Fig. 10 for ms3. Nevertheless, it is interesting to notice\nthat, in contrast to the strongly hysteretic behavior of the mode l I at the transition between\nthe two ferrimagnetic phases, no apparent hysteresis can be obs erved in the present model.\nDiscontinuous character of the transition reflected in bimodality of the relevant observables,\nsuch as the internal energy, disappears at higher temperatures but is still evident at temper-\natures slightly above kBT/|J|= 0.2. In Fig. 11(a) it is demonstrated for D/|J|=−1.48 and\nthe temperatures kBT/|J| ≈0.215 tuned by the reweighing technique for each Lto achieve\ndistributions with the two modes of about the same heights. The inse t gives us some idea\nabout the characteristic tunneling times between the coexisting ph ases forL= 48. However,\nsimilar tothesituationinthemodelIpresented inFig.6, for D/|J|=−1.48thespecific heat\nand staggered susceptibility maxima do not scale with volume. Such a s caling is observed\nonly at slightly lower temperatures for D/|J|=−1.4825 (at least for L≥72), as shown in\nFig. 11(b). At higher temperatures the second-order phase tra nsition with standard Ising\n17−0.02−0.015 −0.01−0.005 00.005 0.0100.511.522.53x 105\ne/|J| \nL=48\nL=72\nL=96\nL=1208 9−505x 10−3\n105 MCSe/|J|(a)\n3 3.5 4 4.5 5−20246810\nln(L)ln(cmax) ln( χs3,max)\n \nD/|J|=−1.47\nD/|J|=−1.48D/|J|=−1.47D/|J|=−1.48\nD/|J|=−1.4825, α/ν=2(b)\nD/|J|=−1.4825, γ/ν=2\n3 3.5 4 4.5 50123456\nln(L)ln(χM\ns3,max), ln(D1M\ns3,max), ln(D2M\ns3,max)\n \nln(χM\ns3,max), γI/νI=1.75\nln(D1M\ns3,max), 1/νI=1.00ln(D2M\ns3,max), 1/νI=1.00(c)\nFIG. 11: (Color online) (a) Energy distributions for D/|J|=−1.48 and different L. The respective\ntemperatures are tuned by the reweighing technique to achie ve approximately equal peak heights.\nTime evolution of the internal energy shown in the inset demo nstrates tunneling between the\ncoexisting phases for L= 48. (b) FSS analysis of the susceptibility χMs3and the specific heat\nc, forD/|J|=−1.47,−1.48 and−1.4825. For D/|J|=−1.4825 the log-log plots are fitted to\nthe exponent 2, corresponding to the system volume. (c) FSS a nalysis of χMs3,D1Ms3,D2Ms3for\nD/|J|=−1.46withthefittedIsingvaluesofthecritical exponentsrati osγI/νI= 1.75,1/νI= 1.00.\ncritical exponents is recovered for D/|J|=−1.46, although again larger system sizes are\nrequired to reach the linear asymptotic regime (Fig. 11(c)). Thus, the tricritical point in\nthe model II is roughly located at ( Dt/|J|,kBTt/|J|) = (−1.47±0.01,0.27±0.04).\nThe above results can be summarized into the phase diagram shown in Fig. 12. As in\n18−1.5−1−0.500.511.5200.20.40.60.811.21.4\nD/|J|kBTc/|J|P\nTCPFRII\nfirst order\nFIG. 12: (Color online) Phase diagram of the model II in ( kBT/|J|−D/|J|) parameter space. The\nempty circles represent the phase transition temperatures kBTc/|J|between the paramagnetic P\nand the ferrimagnetic phase FRII(0,±1,∓1) estimated from the specific heat peaks for L= 48,\nthe filled circle shows a more precise value obtained from the FSS analysis and Binder cumulant\ncrossing, the filled diamond is the tricritical point and the filled squares represent a first-order\ntransition point determined from FSS analysis at D/|J|=−1.48 and the exact value of the GS\ntransition point at Dc/|J|=−3/2. The empty triangles mark the first-order transition relat ed\ndiscontinuities in the order-parameter ms3in theD/|J|increasing ( ⊲) and decreasing ( ⊳) processes.\nthe phase diagram of the model I, the empty circles represent the pseudo-critical points\ndetermined from the specific heat maxima at L= 48 and the empty triangles mark the first-\norder phase transitions located from the jumps in the order param eter loops obtained by\nincreasing and decreasing of the single-ion parameter D/|J|. The filled circle at D/|J|= 0,\nthe square at D/|J|=−1.48 and the filled diamond represent respectively second-order,\nfirst-order and tricritical points, obtained from the FSS analysis. As in Fig. 8, the filled\nsquare at ( Dc/|J|,kBTc/|J|) = (−3/2,0) represents the exact value of the ground-state\nphase transition point.\nIV. CONCLUSIONS\nWe have studied the mixed spin-1/2 and spin-1 Ising ferrimagnets on a triangular lattice\nwithsublattices A, BandC,intwo mixing modes: ( SA,SB,SC) = (1/2,1/2,1)(model I)and\n19(SA,SB,SC) = (1/2,1,1)(model II). The pure spin-1/2 and spin-1 Ising antiferromagne ts on\na triangular lattice show respectively no LRO45and partial LRO for some range of a single-\nion anisotropy parameter at low temperature with quasi-LRO of the Berezinskii-Kosterlitz-\nThouless type at higher temperatures46. In comparison with these models in the present\nferrimagnetic models the frustration is partially accommodated by d ifferent ferrimagnetic\nspin arrangements and thus their critical behavior is rather differe nt from the pure systems.\nOn the other hand, the net magnetization of both ferrimagnetic mo dels is always zero and\nthus they canshow no compensation points. Therefore, fromthis point ofview, the behavior\nof the present mixed-spin models is typical for antiferromagnets r ather than ferrimagnets.\nAs for the critical properties, the model I shows two ferrimagnet ic phases FRI\n1:\n(±1/2,±1/2,∓1) and FRI\n2: (±1/2,∓1/2,0), which may be compared with the ferromag-\nnetic case displaying two ferromagnetic phases ( ±1/2,±1/2,±1) and (±1/2,±1/2,0)29. We\nnote that on bipartite lattices the thermodynamic behavior of the s ystems with ferrimag-\nnetic and ferromagnetic interactions is the same and thus the phas e diagrams would be\nidentical. However, there are substantial differences between th e two phase diagrams for the\ntriangular lattice ferromagnetic and ferrimagnetic models. First of all, due to frustration\nthe transition temperatures for the ferrimagnetic case are signifi cantly reduced and collapse\nwith the ferromagnetic boundary only in the large negative D/|J|limit, when the partial\nfrustration inducing magnetic states on the C-sublattice are comp letely suppressed and the\ncritical temperature for either case tends to the exact spin-1/2 Ising value on a honeycomb\nlatticekBTc/|J|= 0.379747. The frustration is further increased close to the boundary be-\ntween the two ferrimagnetic phases FRI\n1and FRI\n2, which is reflected in the depression in the\norder-disorder phase boundary that is absent in the ferromagne tic model. Nevertheless, the\nfrustration did not seem to affect the standard Ising values of the critical exponents. We\nnote that besides the above presented points of D/|J|=−2 and 0, we also performed the\nFSS analysis in this region of an increased frustration at the P-FRI\n1branch of the phase\ndiagram for D/|J|=−1.6 (not shown) but did not find any deviation larger than statis-\ntical errors from the standard values. However, the most consp icuous difference from the\nferromagnetic case is the presence of the strongly discontinuous phase transition between\nthe ferrimagnetic phases ( ±1/2,±1/2,∓1) and (±1/2,∓1/2,0), which is completely absent\nbetween the ferromagnetic phases ( ±1/2,±1/2,±1) and (±1/2,±1/2,0). This can be ex-\nplained by the fact that in the ferrimagnetic case dramatic changes occur when the two\n20spin-1/2 sublattices A and B, forming a connected honeycomb back bone, switch their mag-\nnetizations between 1 /2 and−1/2 and the spin-1 sublattice C between magnetic ±1 and\nnonmagnetic 0 states. On the other hand, in the ferromagnetic ca se nothing happens in\nsublattices A and B and in sublattice C the change between magnetic ±1 and nonmagnetic\n0 states occurs only gradually in the isolated (mutually directly nonint eracting) spins.\nThe model II has been shown to display only one ordered ferrimagne tic state FRII:\n(0,±1,∓1) with no LRO on sublattice A and an antiferromagnetic LRO on the re maining\nsublattices B and C forming a honeycomb lattice. Similar phase occurs in the pure spin-\n1 triangular antiferromagnet for the single-ion anisotropy parame ter−3/2< D/|J|<0,\nhowever, it is destabilized for D/|J|>046. On the other hand, in the present model II\nthe increasing D/|J|stabilizes the ferrimagnetic phase FRIIand the critical temperature for\nD/|J| → ∞tends to kBTc/|J|= 1.518847, i.e., the exact value of the spin-1/2 Ising model\non a honeycomb lattice when the spin states ±1 are considered in the Hamiltonian instead\nof±1/2. In comparison with the ferromagnetic model II, again the critica l temperatures\nare lowered due to frustration but qualitatively the phase diagrams look similar. As the\nparameter D/|J|is decreased both models show change of the phase transition natu re from\nsecond to first order at a tricritical point. Nevertheless, interes tingly, the strong hysteretic\nbehavior observed intheferromagneticmodel accompanying thefi rst-orderphase transitions\nis not evidenced in the ferrimagnetic one.\nAcknowledgments\nThis work was supported by the Scientific Grant Agency of Ministry o f Education of\nSlovak Republic (Grant Nos. 1/0234/12 and 1/0331/15). The auth ors acknowledge the\nfinancialsupportbytheERDFEU(EuropeanUnionEuropeanRegion alDevelopmentFund)\ngrant provided under the contract No. ITMS26220120047 (activ ity 3.2.).\n∗Electronic address: milan.zukovic@upjs.sk\n1L.L. Goncalves, Phys. Scripta 32, 248 (1985).\n2A. Lipowski, T. Horiguchi, J. Phys. A: Math. Gen. 28, L261 (1995).\n3M. Jaˇ sˇ cur, Physica A 252, 217 (1998).\n214A. Dakhama, Physica A 252, 225 (1998).\n5M. Jaˇ sˇ cur, J. Streˇ cka, Condens. Matter Phys. 8, 869 (2005).\n6T. Kaneyoshi, J.C. Chen, J. Magn. Magn. Mater. 98, 201 (1991).\n7O.F. Abubrig, D. Horv´ ath, A. Bob´ ak, M. Jaˇ sˇ cur, Physica A 296, 437 (2001).\n8T. Kaneyoshi, J. Phys. Soc. Jpn. 56, 2675 (1987).\n9A. Bob´ ak, M. Jurˇ ciˇ sin, Physica A 240, 647 (1997).\n10A. Bob´ ak, Physica A 258, 140 (1998).\n11T. Kaneyoshi, Y. Nakamura, J. Phys.: Condens. Matter 10, 3003 (1998).\n12T. Kaneyoshi, Y. Nakamura, S. Shin, J. Phys.: Condens. Matte r10, 7025 (1998).\n13A. Bob´ ak, Physica A 286, 531 (2000).\n14A. Bob´ ak, O.F. Abubrig, D. Horv´ ath, J. Magn. Magn. Mater. 246, 177 (2002).\n15G.M. Zhang, C.Z. Yang, Phys. Rev. B 48, 9452 (1993).\n16G.M. Buendia, M. Novotny, J. Phys.: Condens. Matter 9, 5951 (1997).\n17W. Selke, J. Oitmaa, J. Phys.: Condens. Matter 22, 076004 (2010).\n18Y. Nakamura, J. Phys.: Condens. Matter 12, 4067 (2000).\n19Y. Nakamura, J.W. Tucker, IEEE Trans. Magn. 38, 2406 (2002).\n20J. Oitmaa, W.-H. Zheng, Physica A 328, 185 (2003).\n21M. Godoy, W. Figueiredo, Physica A 339, 392 (2004).\n22W. Selke, J. Oitmaa, J. Phys.: Condens. Matter 22, 076004 (2010).\n23T. Iwashita, N. Uryu, J. Phys. Soc. Japan 53, 721 (1984).\n24H.F. Verona de Resende, F.C. S´ aBarreto, J.A. Plascak, Phys ica A149, 606 (1988).\n25J.W. Tucker, J. Magn. Magn. Mater. 237, 215 (2001).\n26M. Godoy, V.S. Leite, W. Figueiredo, Phys. Rev. B 69, 054428 (2004).\n27J. Oitmaa, Phys. Rev. B 72, 224404 (2005).\n28J. Oitmaa, I.G. Enting, J. Phys.: Condens. Matter 18, 10931 (2006).\n29M.ˇZukoviˇ c, A. Bob´ ak, arXiv:1412.5811 [cond-mat.stat-mec h].\n30P.E. Theodorakis, N.G. Fytas, Phys. Rev. E 86, 011140 (2012).\n31A.M. Ferrenberg, R.H. Swendsen, Phys. Rev. Lett. 61, 2635 (1988).\n32U. Wolff, Computer Physics Communications 156 (2004) 143.\n33D. York, N. Evensen, M. Martinez, J. Delgado, Am. J. Phys. 72367 (2004).\n34H. Theil, Economic Forecasts and Policy, Vol. XV of Contributions to Ec onomic Analysis ,\n22(North-Holland, Amsterdam, 1961).\n35A.M. Ferrenberg, D.P. Landau, Phys. Rev. B 44, 5081 (1991).\n36D.P. Landau, K. Binder, A Guide to Monte Carlo Methods in Statistical Physics , (Cambridge\nU. Press, Cambridge, 2000).\n37G. Kamieniarz, H.W.J. Bl¨ ote, J. Phys. A: Math. Gen. 26, 201 (1993).\n38K. Binder, Z. Phys. B 43, 119 (1981).\n39D.P. Landau, R.H. Swendsen, Phys. Rev. Lett. 46, 1437 (1981).\n40S. Jin, A. Sen, W. Guo, A.W. Sandvik, Phys. Rev. B 87, 144406 (2013).\n41M.P.M. den Nijs, J. Phys. A: Math. Gen. 12, 1857 (1979).\n42B. Nienhuis, A.N. Berker, E.K. Riedel, M. Schick, Phys. Rev. Lett.43, 737 (1979).\n43B. A. Berg and T. Neuhaus, Phys. Lett. B 267, 249 (1991); Phys. Rev. Lett. 68, 9 (1992).\n44J. Zierenberg, N.G. Fytas, W. Janke, Phys. Rev. E 91, 032126 (2015).\n45G.H. Wannier, Phys. Rev. 79, 357 (1950).\n46M.ˇZukoviˇ c, A. Bob´ ak, Phys. Rev. E 87, 032121 (2013).\n47M.E. Fisher, Rep. Prog. Phys. 30, 615 (1967).\n48Hence, in fact these are only pseudo-critical points\n23" }, { "title": "2011.02001v1.An_attempt_to_simulate_laser_induced_all_optical_spin_switching_in_a_crystalline_ferrimagnet.pdf", "content": "arXiv:2011.02001v1 [cond-mat.mtrl-sci] 3 Nov 2020An attempt to simulate laser-induced all-optical spin swit ching in a crystalline\nferrimagnet\nG. P. Zhang∗, Robert Meadows, and Antonio Tamayo\nDepartment of Physics, Indiana State University, Terre Hau te, IN 47809, USA†\nY. H. Bai\nOffice of Information Technology, Indiana State University, Terre Haute, Indiana 47809, USA\nThomas F. George\nDepartments of Chemistry & Biochemistry and Physics & Astro nomy\nUniversity of Missouri-St. Louis, St. Louis, MO 63121, USA\n(Dated: November 5, 2020)\nInterest in all-optical spin switching (AOS) is growing rap idly. The recent discovery of AOS in\nMn2RuGa provides a much needed clean case of crystalline ferrim agnets for theoretical simulations.\nHere, we attempt to simulate it using the state-of-the-art fi rst-principles method combined with the\nHeisenberg exchange model. We first compute the spin moments at two inequivalent manganese\nsites and then feed them into our model Hamiltonian. We emplo y an ultrafast laser pulse to switch\nthe spins. We find that there is a similar optimal laser field am plitude to switch spins. However, we\nfind that the exchange interaction has a significant effect on t he system switchability. Weakening\nthe exchange interaction could make the system unswitchabl e. This provides a crucial insight into\nthe switching mechanism in ferrimagnets.\nPACS numbers: 75.78.Jp, 75.40.Gb, 78.20.Ls, 75.70.-i\nCentral to the magnetic storage device is the writ-\ning/reading speed of magnetic bits in a storage medium.\nTraditionally, these operations are mostly driven by an\nexternal magnetic field. A full-optical driven spin manip-\nulation could break the speed barrier of several hundred\npicoseconds set by the Zeeman interaction and magnetic\ndipole-dipole interaction. In 1996, Beaurepaire et al.[1]\nshowed that when they shone a 60-fs pulse on the ferro-\nmagnetic nickel thin film, they found a sharp decrease in\nthe Kerr signal within 1 ps. This finding received imme-\ndiate attention worldwide, and a new research field, fem-\ntomagnetism, wasborn[2,3]. Researchintensified, andis\nfar beyond the scope of the original research interest. In\n2007, Stanciu et al.[4] showed that a left-circularly po-\nlarized laser pulse can switch an up-spin to down, while a\nright-circularly polarized laser pulse can switch a down-\nspinup. Thisremarkablepropertyrepresentsaninterest-\ningnew magnetic phenomenononan ultrafasttime scale,\nalthough their compound, GdFeCo, is not new. GdFeCo\nhas been used in traditional magneto-optical recording\n(see the references cited in [5, 6]). However, being to\nable to switch spins on a picosecond time scale optically\nis new, and has raised the possibility for a real applica-\ntion. However,itisunclearhowthelaserpulsecanswitch\nspins directly. Ostler et al.[7] further showed that if the\nlaser intensity is increased above a certain level, regard-\nless of laser helicity, each pulse can flip spins from one\ndirection to another deterministically. They argued that\n∗Author to whom correspondence should be addressed.\n†Electronic address: guo-ping.zhang@outlook.com.there is a threshold intensity that one has to exceed to\nchange from all-optical helicity dependent spin switch-\ning to all-optical helicity independent switching, but the\nactual picture is more complicated [8, 9].\nFor a long time, GdFeCo was the only material that\nshows AOS. Soon, many more materials were found [10–\n14]. However, these materials are mostly amorphous,\nwhich introduces an uncertainty in theoretical simula-\ntions and represents a formidable task. In 2017, Vomir\net al.[15] reported the first observation of AOS in a\nPt/Co/Pt ferromagnetic stack, but the switching is not\ncomplete. Very recently, Banerjee et al.[16] showed\nsingle-pulse all-optical toggle switching of magnetization\nin Mn2RuGa. Mn 2RuGa is a ferrimagnetic Heusler com-\npound, with a cubic structure. Two Mn atoms are not\nequivalent, and have different spin moments. They are\nantiferromagnetically coupled. As shown before [17], fer-\nrimagnets have a big advantage over ferromagnets and\nantiferromagnets. This offers an ideal theoretical model.\nIn this paper, we investigate all-optical switching in\nMn2RuGa. Different from prior studies, we compute the\nspin moments at two Mn sites using the first-principles\ndensity functional theory. These spin moments are fed\ninto the Heisenberg exchange model with both spin-orbit\ncoupling and a harmonic potential [18, 19]. We find that\nnot any arbitrary laser field amplitude can switch spins.\nThere is a narrowwindow of opportunity where the spins\nat two Mn sites can be switched into their respective op-\nposite directions. Because of the strong spin moments\nat two Mn sites, its switching is very stable. Quite dif-\nferent from other systems, we find that if we reduce the\nexchange interaction, the spins precess strongly at both\nMn sites, very much like a regular antiferromagnet, in-2\nstead of a ferrimagnet. These strong spin oscillation oc-\ncur at both Mn 1and Mn 2sites. Their oscillation period\nis inverselyproportionalto the laser field amplitude, sim-\nilar to the Rabi frequency in a two-level system. For a\nweak exchange interaction case, regardless of the magni-\ntude ofthe laserfield amplitude, the spin switchingis not\nobserved. This points out an entirely different scenario\nfrom ferromagnetic cases [19] and demagnetization [20].\nThe rich picture that is found here revealsa crucial effect\nof the effect of exchange interaction on AOS, and should\nmotivate further experimental and theoretical investiga-\ntions in the future.\nMn2RuGa is a Heusler ferrimagnet, with a stoichio-\nmetric composition of X2YZand space group F¯43m.\nTwo Mn atoms, Mn 1and Mn 2, are situated at (4 a)\nand (4c), which are magnetically inequivalent [21, 22].\nTheir spins are antiferromagnetically coupled. To prop-\nerlyinvestigatemagneticpropertiesofMn 2RuGa, weem-\nploythe density functional theoryusing the full-potentialaugmented plane wave method as implemented in the\nWien2k code [23, 24]. Our first-principles calculation\nshows that Mn 1has a spin moment of 3.17232 µBand\nMn2has -2.30765 µB. This agreeswith priorcalculations\n[21, 22]. So each cell has a net spin moment of 1.02394\nµB, close to unity, which is consistent with the nature of\na stoichiometric half-metal [25]. The spin moments on\nRu and Ga are very small, and will be ignored below.\nIt is not often recognized that the large spin mo-\nments on Mn atoms are advantageous since the spin-\norbit torque is proportional to the spin moment [18, 19].\nSinceitisnotpossibletosimulateall-opticalspinreversal\nat the first-principles level, in the following we will feed\nthese two spin moments into our Heisenberg-exchange\ncoupled harmonic model [17, 19, 26, 27], and limit our-\nselves to a small system with 101 lattice sites along the x\naxis and yaxis, respectively, with two monolayers along\nthezaxis. Our Hamiltonian is\nH=/summationdisplay\ni/bracketleftbiggp2\ni\n2m+V(ri)+λLi·Si−eE(r,t)·ri/bracketrightbigg\n−/summationdisplay\nijJexSi·Sj, (1)\nwhere terms from the left to right are respectively the\nkinetic energy operator of the electron, the potential en-\nergyoperator,the spin-orbitcoupling, the interactionbe-\ntween the laser and system, and the exchange interaction\nbetween spins. Our exchange parameter Jexis still time-\nindependent, although prior studies have shown that the\nexchange interaction itself could be affected by the elec-\ntricfield[28,29] .λisthe spin-orbitcouplingconstant, Li\nandSiare the orbital and spin angular momenta at site\ni, respectively, and pandrare the momentum and posi-\ntion operators of the electron, respectively. We choose a\nspherical harmonic potential V(ri) =1\n2mΩ2r2\niwith sys-\ntem frequency Ω. E(r,t) is the laser field. This model is\nthe only magnetic field-free model currently available to\nsimulate spin reversal, while the commonly used model\nemploys an effective magnetic field [7], which should be\navoided. It represents a small step towards a complete\nmodel.\nIn order to compute the spin change, we solve the\nHeisenberg equation of motion [20] for each spin oper-\nator at every site under laser excitation [17]. Figure 1(a)\nshows the spin zcomponent at two Mn sites as a func-\ntion of time. We employ a laser pulse of 60 fs, with a\nfield amplitude of 0.017 V /˚A. We see that the spin at\nthe Mn 1site starts from the positive zaxis (see the cir-\ncles). Upon laser excitation, it switches over the negative\nzaxis, while the spin at the Mn 2site switches up from\nits−zdirection. This is consistent with the experimen-\ntal observation [16]. The strong spin moment stabilizes\nthe entire switching process. However, not any arbitrarylaser field amplitude can lead to faithful switching. Fig-\nure 1(b) shows how the final spin changes with the laser\nfield amplitude. The dependence is highly nonlinear. If\nwe use a weak laser pulse, there is little change in spins\nat both sites. But if the laser field is too strong, the\nspins overturn toward the xyplane, so there is no spin\nreversal either. We find that the optimal field amplitude\nis 0.017 V /˚A, whose result is shown in Fig. 1(a). In this\nregard, Mn 2RuGa is pretty much similar to other ferri-\nmagnets where there is an optimal amplitude [17, 19].\nMicroscopically, the real situation is more complicated.\nTo this end, there is no generic understanding of spin\nswitching in both ferromagnets [14, 15] and ferrimagnets\n[4]. It has been often argued that the angular momen-\ntum exchange between two spin sublattices in ferrimag-\nnetic GdFeCo [30] is the key to AOS. Theoretically, this\nmomentum exchange picture is interesting, but such mo-\nmentum exchange between sublattices, if it exists, occurs\nall the time through the exchange interaction, with or\nwithout the laser. In other words, it must be something\nextra due to the laser that switches the spin. In GdFeCo,\narigoroustesting isdifficult because itis amorphous,and\nit is difficult to tell whether a model system really repre-\nsents a true GdFeCo sample. This brings ambiguity to a\ntheoretical simulation. Mn 2RuGa removes this ambigu-\nity completely.\nAs a first test, we investigate the effect of the exchange\ninteraction on AOS. We reduce the exchange interaction\nfrom 0.1 eV to 0.001 eV. From prior studies, we know\nsuch a reduction does not constitute a major issue for3\n0 0.01 0.02 0.03 0.04\nA0(V/Å)−2−1012Sz(h− )−400 −200 0 200 400 600Time (fs)\n−2−1012Sz(h− ) Mn1\nMn2(a)\n(b)\nFIG. 1: (a) The zcomponent ofthe spins at the Mn 1and Mn 2\nsites as a function of time at the optimal laser field amplitud e.\nHere, thelaser amplitudeis0.017 V /˚A, andthepulseduration\nis 60 fs. The empty circles denote the spin at the Mn 1site,\nwhile the boxes refer to the spin at the Mn 2site. We see there\nis a clear spin reversal upon laser excitation. (b) Dependen ce\nof theSzas a function of the laser field amplitude A0. The\nempty circles refer to the spin at Mn 1site, and the boxes refer\nto the spin at Mn 2site. A weak laser field does not reverse\nthe spins, but a too strong laser field can not either. There is\na narrow window that one can switch spins.\n0 0.010.020.030.04\nA0(V/Å)4006008001000Period (fs)\n0 0.010.020.030.04\nA0(V/Å)−1012Sz(h− )−500 0 500 1000 1500 2000Time (fs)\n−2−1012S(h− )Sx\nSy\nSz(a)\n(b) (c)\nFIG. 2: (a) Spin precession at Mn 1site under a reduced\nexchange interaction. Here we choose J= 0.001 eV. The rest\nof parameters are the same as those in Fig. 1. The solid,\ndotted, and dashed lines denote the x,y, andzcomponents\nof the spin, respectively. (b) The spin oscillation period d e-\ncreases with laser field amplitude A0. (c) At any of the laser\nfield amplitudes, spin reversal is not found. Only a strong\noscillation is noticed. The solid line is the time-average o f the\nspin, and the dotted and dashed lines refer to the maximum\nand minimum spin values, respectively. The effect of the ex-\nchange interaction on spin reversal is much more pronounced\nin Mn 2RuGa than in other materials.demagnetization in a system with a small spin moment\n[20]. Figure 2(a) shows the spin change at Mn 1(the sit-\nuation is similar at Mn 2), where the solid, dotted, and\ndashed lines denote the x,y, andzcomponents, respec-\ntively. Both the laser duration and amplitude are exactly\nthe same as those in Fig. 1. We see that there is a strong\noscillation in all these three components. We note in\npassing that these three components must obey the op-\nerator permutation [8], where they can not be considered\na linear reversal [31]. In principle, we need to cut off the\nsimulation around 1 ps, after which we need to introduce\ndamping, but to demonstrate the high accuracy of our\ncalculation, we do not use the damping. These strong\noscillations resemble a pure antiferromagnetic case. The\nspins at two neighboring sites are out of phase and re-\nmain antiferromagnetically coupled, even upon laser ex-\ncitation. The laser pulse essentially initiates the spin dy-\nnamics, and the exchange interaction takes over, without\nswitching the spins. For this reason, the angular momen-\ntum exchange picture for AOS can not explain this even\nin the same ferrimagnet. The period of the oscillation\nis not determined by the exchange interaction and spin\nmoment alone. Figure 2(b) shows that as we increase the\nlaser field amplitude, the period becomes shorter. The\nsmall fluctuation at the largest amplitudes is due to the\nperiodsamplingbecausetheoscillationisnotstrictlyhar-\nmonic. This laser-field dependence of the oscillation pe-\nriod is very similar to the Rabi period dependence. For\nall the field amplitudes that we investigate, we do not\nsee a case where the spins are reversed. Figure 2(c) illus-\ntrates the average (solid line), maximum (dotted line),\nand minimum (dashed line) of the final spin. We see the\naverage spin never becomes negative (the initial spin is\nalong the + zaxis). The maximum and minimum values\nshow the limits of spin. Our results point out an impor-\ntant fact: In a ferrimagnet, the effect of the exchange\ninteraction is far more complicated than thought.\nNow, we have two cases: One shows AOS, and the\nother does not. We can directly check whether the prior\ncriteria proposed by Mentink et al.[30] apply to them.\nTheir argument is based on a two-spin system, so for the\npure exchange interaction, the spins at two sublattices\nmust obey the scalar form of spins, ∂S1/∂t=−∂S2/∂t,\nwith the extra term from demagnetization. In our sys-\ntem, each spin is coupled with more than four neighbor-\ningspins, sowetaketwoneighboringspinsasanexample.\nFor the above nonswitchable case (Fig. 2), we find that\nS1x+S2xandS1y+S2yare not constant, so they do not\nobey∂S1/∂t=−∂S2/∂t. OurS1x+S2xdecreasesfrom 0\nto about −1¯hwith oscillations, while S1y+S2yincreases\nfrom 0 to about +1¯ hat the same rate. For our switch-\nable case (Fig. 1), ∂S1/∂t=−∂S2/∂tis not fulfilled\neither. Instead, we find that our result obeys the vector\nformS1(t)/|S1(0)|=−S2(t)/|S2(0)|. This shows that\nthe simple argument based on a two-spin model is not\napplicable to our realistic case. We plan to investigate\nthis issue further in a much larger system.\nIn conclusion, we have carried out a joint first-4\nprinciples density functional theory and model simula-\ntion of all-optical spin reversal in Mn 2RuGa. We are\nable to find a case that the spins can be switched with-\nout employing a magnetic field. The system also shows\nan optimal laser electric field amplitude, with the same\nprofile like those in other systems. The spins at Mn 1site\nare switched from the + zaxis to−zaxis, while those at\nMn2site are switched from the −zaxis to + zaxis. This\nisfullyconsistentwiththeexperimentalfindings[16]. We\nfind that the exchange interaction has a significant effect\non the switching. When we reduce the exchange to 0.001\neV, we find the system becomes unswitchable. It behaves\nlike a regular antiferromagnet. Our finding is expected\nto motivate further theoretical and experimental investi-\ngations in the future.\nAcknowledgments\nThis work was solely supported by the U.S. De-\npartment of Energy under Contract No. DE-FG02-06ER46304. Part of the work was done on Indiana\nState University’s Quantum Cluster and High Perfor-\nmance computers. This research used resources of the\nNational Energy Research Scientific Computing Center,\nwhich is supported by the Office of Science of the U.S.\nDepartment of Energy under Contract No. DE-AC02-\n05CH11231. Our calculations also used resources of the\nArgonne Leadership Computing Facility at Argonne Na-\ntional Laboratory, which is supported by the Office of\nScience of the U.S. Department of Energy under Con-\ntract No. DE-AC02-06CH11357.\nAvailability of data. The data that support the find-\nings of this study are available from the corresponding\nauthor upon reasonable request.\nhttps://orcid.org/0000-0002-1792-2701\n[1] E. Beaurepaire, J. C. Merle, A. Daunois, and J.-Y. Bigot,\nUltrafast spin dynamics in ferromagnetic nickel, Phys.\nRev. Lett. 76, 4250 (1996).\n[2] G. P. Zhang, W. H¨ ubner, E. Beaurepaire, and J.-Y.\nBigot, Laser-induced ultrafast demagnetization: Femto-\nmagnetism, A new frontier? Topics Appl. Phys. 83, 245\n(2002).\n[3] A. Kirilyuk, A. V. Kimel, and Th. Rasing, Ultrafast op-\ntical manipulation of magnetic order, Rev. Mod. Phys.\n82, 2731 (2010). Erratum: Rev. Mod. Phys. 88, 039904\n(2016).\n[4] C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk, A.\nTsukamoto, A. Itoh, and Th. Rasing, All-optical mag-\nnetic recording with circularly polarized light, Phys.\nRev. Lett. 99, 047601 (2007).\n[5] G. P. Zhang, T. Latta, Z. Babyak, Y. H. Bai, and T.\nF. George, All-optical spin switching: A new frontier in\nfemtomagnetism – A short review and a simple theory,\nMod. Phys. Lett. B 30, 1630005 (2016).\n[6] G. P. Zhang, M. Murakami, M. S. Si, Y. H. Bai, and\nT. F. George, Understanding all-optical spin switching:\nComparison between experiment and theory, Mod. Phys.\nLett. B32, 1830003 (2018).\n[7] T. A. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell,\nU.Atxitia, O.Chubykalo-Fesenko,S.ElMoussaoui, L.Le\nGuyader, E. Mengotti, L. J. Heyderman, F. Nolting, A.\nTsukamoto, A. Itoh, D. Afanasiev, B. A. Ivanov, A. M.\nKalashnikova, K. Vahaplar, J. Mentink, A. Kirilyuk, Th.\nRasing, and A. V. Kimel, Ultrafast heating as a sufficient\nstimulusfor magnetization reversal inaferrimagnet, Nat.\nCommun. 3, 666 (2012).\n[8] G. P. Zhang, Microscopic theory of ultrafast spin linear\nreversal, J. Phys.: Condens. Mattter 23, 206005 (2011).\n[9] G. P. Zhang and T. F. George, Thermal or nonther-\nmal? That is the question for ultrafast spin switching in\nGdFeCo, J. Phys.: Condens. Matter 25, 366002 (2013).[10] S. Mangin, M. Gottwald, C-H. Lambert, D. Steil, V. Uh-\nlir, L. Pang, M. Hehn, S. Alebrand, M. Cinchetti, G.\nMalinowski, Y. Fainman, M. Aeschlimann, and E. E.\nFullerton, Engineered materials for all-optical helicity -\ndependent magnetic switching, Nat. Mater. 13, 286\n(2014).\n[11] A. Hassdenteufel, C. Schubert, B. Hebler, H. Schulthei ss,\nJ. Fassbender, M. Albrecht, and R. Bratschitsch, All-\noptical helicity dependent magnetic switching in Tb-Fe\nthin films with a MHz laser oscillator, Opt. Express 22,\n10017 (2014).\n[12] C. Schubert, A. Hassdenteufel, P. Matthes, J. Schmidt,\nM. Helm, R. Bratschitsch, and M. Albrecht, All-optical\nhelicitydependentmagnetic switchinginanartificial zero\nmoment magnet, Appl. Phys. Lett. 104, 082406 (2014).\n[13] S. Alebrand, U. Bierbrauer, M. Hehn, M. Gottwald,\nO. Schmitt, D. Steil, E. E. Fullerton, S. Mangin, M.\nCinchetti, and M. Aeschlimann, Subpicosecond magne-\ntization dynamics in TbCo alloys, Phys. Rev. B 89,\n144404 (2014).\n[14] C.-H. Lambert, S. Mangin, B. S. D. Ch. S. Varaprasad,\nY. K. Takahashi, M. Hehn, M. Cinchetti, G. Malinowski,\nK. Hono, Y. Fainman, M. Aeschlimann, and E. E. Fuller-\nton, All-optical control of ferromagnetic thin films and\nnanostructures, Science 345, 1337 (2014).\n[15] M. Vomir, M. Albrecht, and J.-Y. Bigot, Single shot all\noptical switching of intrinsic micron size magnetic do-\nmains of a Pt/Co/Pt ferromagnetic stack, Appl. Phys.\nLett.111, 242404 (2017).\n[16] C. Banerjee, N. Teichert, K. Siewierska, Z. Gercsi, G.\nAtcheson, P. Stamenov, K. Rode, J. M. D. Coey and\nJ. Besbas, Single pulse all-optical toggle switching of\nmagnetization without Gd: The example of Mn 2RuxGa,\narXiv:1909.05809.\n[17] G. P. Zhang, Z. Babyak, Y. Xue, Y. H. Bai, and T.\nF. George, First-principles and model simulation of all-5\noptical spin reversal, Phys. Rev. B 96, 134407 (2017).\n[18] G. P. Zhang, Y. H. Bai, and T. F. George, A new and\nsimple model for magneto-optics uncovers an unexpected\nspin switching, EPL 112, 27001 (2015).\n[19] G. P. Zhang, Y. H. Bai, and T. F. George, Switching\nferromagnetic spins by an ultrafast laser pulse: Emer-\ngence of giant optical spin-orbit torque, EPL 115, 57003\n(2016).\n[20] G. P. Zhang, M. Murakami, Y. H. Bai, T. F. George, and\nX. S. Wu, Spin-orbit torque-mediated spin-wave excita-\ntion as an alternative paradigm for femtomagnetism, J.\nAppl. Phys. 126, 103906 (2019).\n[21] I. Galanakis, K. ¨Ozdogan, E. Sasioglu and S. Bl¨ ugel,\nEffect of disorder on the magnetic propertie of cu-\nbic Mn 2RuxGa compounds: A first-principles study, J.\nAppl. Phys. 116, 033903 (2014).\n[22] L. Yang, B. Liu, F. Meng, H. Liu, H. Luo, E. Liu, W.\nWang and G. Wu, Magnetic properties of Heusler alloy\nMn2RuGe and Mn 2RuGa ribbons, J. Mag. Magn. Mater.\n379, 1 (2015).\n[23] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka,\nandJ. Luitz, WIEN2k, AnAugmentedPlane Wave+Lo-\ncal Orbitals Program for Calculating Crystal Properties\n(Karlheinz Schwarz, Techn. Universit¨ at Wien, Austria,\n2001).\n[24] G. P. Zhang, Y. H. Bai, and T. F. George, Ultrafast\nreduction of exchange splitting in ferromagnetic nickel,\nJ. Phys.: Condens. Mattter 28, 236004 (2016).\n[25] H. Kurt, K. Rode, P. Stamenov, M. Venkatesan, Y.-C.Lau, E. Fonda, and J. M. D. Coey, Cubic Mn 2Ga thin\nfilms: crossing the spin gap with ruthenium, Phys. Rev.\nLett.112, 027201 (2014).\n[26] M. Murakami, Z. Babyak, M. Giocolo, and G. P. Zhang,\nQuantum mechanical interpretation of the ultrafast all-\noptical spin switching, J. Phys.: Conden. Matter 29,\n184002 (2017).\n[27] G. P. Zhang, Y. H. Bai, and T. F. George, Is perpendic-\nular magnetic anisotropy essential to all-optical ultrafa st\nspin reversal inferromagnets? J. Phys.: Condens. Matter\n29, 425801 (2017).\n[28] J. H. Mentink, K. Balzer, and M. Eckstein, Ultrafast\nand reversible control of the exchange interaction in Mott\ninsulators, Nat. Commun. 6, 6708 (2015).\n[29] R. V. Mikhaylovskiy, E. Hendry, A. Secchi, J. H.\nMentink, M. Eckstein, A. Wu, R. V. Pisarev, V. V.\nKruglyak, M. I. Katsnelson, Th. Rasing, and A. V.\nKimel, Ultrafast optical modification of exchange inter-\nactions in iron oxides, Nat. Commun. 6, 8190 (2015).\n[30] J. H. Mentink, J. Hellsvik, D. V. Afanasiev, B. A. Ivanov ,\nA. Kirilyuk, A. V. Kimel, O. Eriksson, M. I. Katsnelson,\nand Th. Rasing, Ultrafast spin dynamics in multisublat-\ntice magnets, Phys. Rev. Lett. 108, 057202 (2012).\n[31] K. Vahaplar, A. M. Kalashnikova, A. V. Kimel, D.\nHinzke, U. Nowak, R. Chantrell, A. Tsukamoto, A. Itoh,\nA. Kirilyuk, and Th. Rasing, Ultrafast path for opti-\ncal magnetization reversal via a strongly nonequilibrium\nstate, Phys. Rev. Lett. 103, 117201 (2009)." }, { "title": "1610.09200v1.Spin_Orbit_Torque_Efficiency_in_Compensated_Ferrimagnetic_Cobalt_Terbium_Alloys.pdf", "content": "1 \n Spin-Orbit Torque Efficiency in Compensated Ferrimagnetic Cobalt -Terbium Alloys \nJoseph Finley1 and Luqiao Liu1 \n1Department of Electrical Engineering and Computer Science, \nMassachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA \n \n Despite the potential advantag es of information storage in antiferromagnetically coupled \nmaterials, it remains unclear whether one can control the magnetic moment orientation efficiently because \nof the cancelled magnetic moment. Here, we report spin -orbit torque induced magnetization switching of \nferrimagnetic Co 1-xTbx films with perpendicular magnetic anisotropy. Current induced switching is \ndemonstrated in all of the studied film compositions, including those near the magnetization \ncompensation point . The spin-orbit torque induced effective field is further quantified in the domain wall \nmotion regime . A divergent behavior that scales with the inverse of magnetic moment is confirmed close \nto the compensation point, which is consistent with angular momentum conservation . Moreover , we also \nquantify the Dzyaloshinskii -Moriya interaction energy in the Ta/ Co1-xTbx system and we find that the \nenergy density increases as a function of the Tb concentration . The demonstrated spin -orbit torque \nswitching, in combinatio n with the fast magnetic dynamics and minimal net magnetization of \nferrimagnetic alloys, promises spintronic devices that are faster and with higher density than traditional \nferromagnetic systems. \n \n 2 \n I. Introduction \n There has been great interest recently in using antiferromagnet ically coupled materials as opposed \nto ferromagnet ic materials (FM) to store information . Compared with FM, antiferromagnet ically coupled \nsystems exhibit fast dynami cs, as well as immunities against perturbation s from external magnetic field s, \npotentially ena bling spintronic devices with higher speed and density [1,2] . Rare earth (RE) – transition \nmetal (TM) ferrimagnetic alloys are one potential candidate material for realizing such devic es. Inside \nRE-TM alloys, the moments o f TM elements (such as Fe, Co, Ni) and the RE elements (e.g., Gd, Tb , Ho, \netc) can be aligned with anti -paralle l orientations due to the exchange interaction between the f and d \nelectrons [3]. By varying the relative concentrations of the two species, one can reach compensation \npoints where the net magnet ic moment or angular momentum goes to zero [4–7]. Moreover, because of \nthe different origins of magnetism in the two species, transport related properties are dominated by TM in \nthese alloys, providing a way to read out the magnetic state even in a compensated system. In this work, \nwe show that by uti lizing the current induced spin -orbit torque, one can switch magnetic moments in \nTa/Co 1-xTbx bilayer films. Particularly, we found that effective fields generated from the spin -orbit torque \nscaled with the inverse of magnetization and reached maximum when the composition approaches the \nmagnetic compensation point. The large effective spin -orbit torque and the previously demonstrated fast \ndynamics [8,9] in these ferrimagnetic systems provide a prom ising platform for high speed spintronic \napplications. \n \nII. Characterization of Magnetic Properties \n Spin-orbit torque (SOT) originates from spin -orbit interaction (SOI) induced spin generation in \nthe bulk (i.e., the spin Hall effect) [10,11] or the surface (i.e., the Rashba -Edelstein effect) [12,13] of solid \nmaterials. So far , SOTs have proven to be an efficient method of controlling the ferromagnetic state of \nnanoscale devices [14–17]. Recently, it was demonstrated that one can utilize SOT to switch TM-\ndominant Co FeTb ferrimagnet s with perpendicular magnetic anisotropy (PMA) [18]. It is therefore \ninteresting to ask what the relationship is between the chemical composition of RE -TM alloys and the 3 \n SOT effi ciency, and whether or not one can switch compensated ferrimagnets using SOT. To answer \nthese questions, we grew a series of Ta(5)/Co 1-xTbx(t)/Ru(2 ) (thickness in nm) films using magnetron \nsputtering. The Co1-xTbx alloys were deposited by co -sputtering Co and Tb sources with different \nsputtering powers. The concentration of Tb x was calculated from the deposition rates and varied between \n0.1 and 0.3 , while the layer thickness t ranges from 1.7 nm to 2.6 nm. The magnetic properties of the \ndeposited films were examined using vibra ting sample magnetometry (VSM), and PMA was observed for \nall samples [Fig. 1 (a)]. Furthermore, the magnetic moment goes through zero and the coercive fields \nreach their maximum around x ≈ 0.22, which is consistent with the room temperature magnetic moment \ncompensation point xcM reported earlier [6,7] . The dependence of the magnetic moment on the Tb \nconcentration is summarized in Fig. 1(c), which agrees well with t he trend line calculated by assuming an \nanti-parallel alignment between the Co and Tb moment (dashed lines). The samples were then patterned \ninto Hall b ars with dimensions 4 x 44 μm2 [Fig. 1(d)]. The anomalous Hall resistance ( RAH) vs magnetic \nfield H curves measured from these samples are plotted in Fig. 1(b) and summarized in Fig. 1(e). The \npolarity of the hysteresis loops changes sign across xcM, consistent with previous studies where RAH is \nshown to be dominated by the momentum of the Co sublattice [19,20] . \n \nIII. Current -Induced Switching \nFig. 2 (a) illustrates the current -induced magnetic switching for the series of samples. During \nthese measurements, in -plane magnetic fields of ±2000 Oe were applied in the current flowing direction \n[y axis in Fig. 1 (d )]. Previous studies showed that an in -plane field is necessary to ensure dete rministic \nmagnetic switching of PMA films , as it can break the symmetry between two equivalent final states [21–\n26] . As shown in Fig. 2(a ), the current -induced switching shows opposite p olarities under the positive \nand negative applied fields, consistent with the model of SOT induced switching [26]. Furthermore, under \nthe same in -plane field, the switching polarity changes sign as the samples go from being Co -dominant to \nTb-dominant . This phenomen on can be explained by considering a macro spin model as shown in Fig. \n2(b). In SOT switching, the Slonczewski torque [27] is proportional to 𝒎̂×(𝝈̂×𝒎̂), where 𝒎̂ is the unit 4 \n vector along the magnetic moment direction and 𝝈̂ is the orientation of electron spins generated from SOI \n[along the 𝒙̂ direction in Fig. 1(d )]. Because the torque is an even function of the local magnetic moment \n𝒎̂, effects from both sublattices in the RE -TM alloy add constructively [1,28] . At equilibrium positions, \n𝒎̂ and 𝝈̂ are perpendicular to each other , and it is usually conve nient to use an effective field [22] 𝐻𝑆𝑇∝\n𝝈̂×𝒎̂ to analyze the SOT effect on magnetic switching. The equilibrium position of 𝒎̂ can then be \ndetermined by balancing the anisotropy field 𝐻𝑎𝑛, the applied in -plane field 𝐻𝑦, and 𝐻𝑆𝑇. As shown in \nFig. 2(b ), when 𝐻𝑦 and 𝝈̂ are given by the illustrated directions, the final orientation of the Co sublattice \nmagnetic moment will be close to the +𝒛̂ direction f or the Co dominant sample and close to the −𝒛̂ \ndirection for the Tb dominant sample, giving rise to opposite Hall voltages. Based upon this analysis, the \ncurrent induced switching should exhibit the same polarity change across the compensation point as the \nmagnetic field induced switching. We note that all of the samples follow this rule except for the \nCo0.77Tb0.23 sample, where the field switching data had determined it to be Tb -dominant but the current \ninduced switching corresponds to a Co -dominant sample. A careful study on this sample reveals that this \nchange simply arises from Joule heating induced temperature change. In RE-TM alloys , the magnetic \nmoments of the RE atom s have stronger temperature dependence compared with TM atoms . \nConsequently , when the temperature increases, a higher RE concentration is ne cessary to achieve the \nsame magnetic moment compensation point [3–9,29] . By measuring the RAH vs H curves of the \nCo0.77Tb0.23 sample under different applied currents , we found that the polarity of the field induced \nswitching did change si gn when the current density is higher than 2 × 107 A∙cm-2, suggest ing that the \nsample underwent a (reversible) transition from Tb -dominant to Co -dominant . \n \nIV. Quantitative Determination of Spin-Orbit -Torque Efficiency \n The critical current of SOT induced switching in a multi -domain sample is influenced by defect -\nrelated factors such as domain nucleation and domain wall (DW) pinning [24]. Therefore, the SOT \nefficiency cannot be simply extracted using the switching current values determined in Fig. 2 ( a). To 5 \n quantify the SOT in our samples, we measured the SOT induced effective field in the DW motion regime \nby comparing it with the applied perpendicular field, using the approach developed by C. F. Pai et \nal. [25]. It has been shown that in PMA films with N éel DWs , the Slonczewski term acts on the DW as an \neffective perpendicular magnetic field and induces DW motion [Fig. 3(g)] [21–25].Therefore, by \nmeasuring the current induced shift in the RAH vs Hz curves, one can determine the magnitude of the SOT. \nFig. 3 (a) and (b) show typical field induced switching curves for a Co-dominant sample Co 0.82Tb0.18 and a \nTb-dominant sample Co0.75Tb0.25. Under the applied current of ±3 mA and in -plane field of 2000 Oe, the \ncenters of hysteresis loops are offset from zero , with opposite values for opposite current directions. The \ncurrent dependence of the offset fiel ds are summarized in Fig. 3 (c ) and (d) for 𝐻𝑦=0 and ±2000 Oe, \nwhere a linear relationship between the offset field and the applied current is obtained. In these plots, the \nratio between the offset field 𝐻𝑧𝑒𝑓𝑓 and current density 𝐽𝑒 curve represents the efficiency of the SOT at the \nTa/RE -TM interface, defined as 𝜒 ≡𝐻𝑧𝑒𝑓𝑓\n𝐽𝑒. 𝜒 as a function of applied 𝐻𝑦 for Co 0.82Tb0.18 and Co 0.75Tb0.25 \nsamples are plotted in Fig. 3 (e ) and (f), respectively . 𝜒 grows linearly in magnitud e for small values of \n𝐻𝑦, until reaching the saturation efficiency 𝜒𝑠𝑎𝑡 at a large in-plane field 𝐻𝑦𝑠𝑎𝑡. The evolution of 𝜒 as a \nfunction of 𝐻𝑦 comes from the chirality change of the DWs in the sample. It is known that because of the \nDzyalosh inskii -Moriya interaction (DMI) mechanism at the heavy metal/magnetic metal interface [30] or \ninside the bulk of RE -TM alloy [31], stable Néel DW with spontaneous chiralities are formed. Under zero \n𝐻𝑦, the DWs do not favor either sw itching polarities, leading to a zero offset field. As 𝐻𝑦 increases, the \nDMI induced effective field 𝐻𝐷𝑀𝐼 is partially canceled, and DWs start to move in directions that facilitate \nmagnetic switching. 𝐻𝑦𝑠𝑎𝑡 therefore represents the minimum field that is required to completely overcom e \n𝐻𝐷𝑀𝐼 and 𝜒𝑠𝑎𝑡 represents the maximum efficiency of the SOT . \n Fig. 4 (a) illustrates the dependence of 𝜒𝑠𝑎𝑡 on x in Co1-xTbx samples. It can be seen that 𝜒𝑠𝑎𝑡 \ndiverges near xCM, with the larges t value occurring for the sample with smallest magnetization. This result \nis consistent with the spin torque theory, where the ratio between the SOT effective field and applied 6 \n charge current is 𝜒𝑠𝑎𝑡=(𝜋/2)(𝜉ℏ/2𝑒𝜇0𝑀𝑠𝑡) [25,32]. Here 𝜉=𝐽𝑆𝐽𝑒⁄ represents the effective spin Hall \nangle, ℏ\n2𝑒𝐽𝑠 is the spin current density, ℏ is Planck’s constant, 𝜇0 is the vacuum permeability , and 𝑀𝑠 is \nthe saturation magnetization. Note that this model of spin torque is based upon the conservation of total \nangular momentum. Previously it has been suggested to utilize ferrimagnetic materials with minimized \n𝑀𝑠 to increase the efficiency of spin torqu e induced switching [33]. However, it was not verified if an \nefficient spin absorption could be achieved at the surface of a ferrimagnet material with antiparallel \naligned sublattices. Moreover, because of the mixture between the spin angular momentum and orbital \nangular momentum in RE -TM alloys, there have been debates over the conservation of total angular \nmomentum in these systems [34]. Our experiment al results provide clear evidence on the strong \nefficiency of spin orbit torques in antiferromagnetically coupled materials. Within the experimenta l \naccuracy, we found that the effective field from the SOT does follow the simple trend given by 1/𝑀𝑆𝑡 \n[dashed lines in Fig. 4(a)], reflecting total angular momentum conservation. 𝜉 in our samples is \ndetermined to be ~0.0 3, smaller than previously repo rted values from Ta/magnetic layer devices, possibly \ndue to the relatively smaller spin-mixing conductance at the Ta/CoTb interface [35]. \nIn addition to the magnetic moment compensation point xcM, inside RE -TM systems there also \nexists an angular momentum compensation point xcJ due to the different g factors associated with spin and \norbit al angular moment um. For our Co 1-xTbx system, using the g factors of Co (~2.2) and Tb (~1.5) \natoms [36,37] , along with the relation 𝐽𝐶𝑜(𝑇𝑏)= 𝑀𝐶𝑜(𝑇𝑏)/𝛾𝐶𝑜(𝑇𝑏), where 𝛾𝐶𝑜(𝑇𝑏)= −𝑔𝐶𝑜(𝑇𝑏)𝜇𝐵/ℏ (𝜇𝐵 \nbeing the Bohr magneton, 𝛾𝐶𝑜(𝑇𝑏) the gyromagnetic ratio, and 𝐽𝐶𝑜(𝑇𝑏) the total angular mo mentum per \nunit volume), we determine xcJ to be ~17%, which is within the range of the studied samples and lower \nthan xcM. Previously it was demonstrated that ultrafast field-driven magnetic dynamics could be excited \naround xcJ [8,9] . According to Landau -Lifshiz -Gilbert equation of a ferr imagnetic system [27,29,38] , the \nspin torque term leads to 𝑑𝒎̂\n𝑑𝑡~−𝛾𝑒𝑓𝑓ℏ𝐽𝑠\n2𝑒𝜇0𝑀𝑆𝑡(𝒎̂×𝝈̂×𝒎̂) , where 𝛾𝑒𝑓𝑓=(𝑀𝐶𝑜−𝑀𝑇𝑏)/(𝐽𝐶𝑜−𝐽𝑇𝑏) is \nthe effective gyromagnetic ratio and 𝑀𝑆=𝑀𝐶𝑜−𝑀𝑇𝑏 [8,9] . When 𝐽𝐶𝑜−𝐽𝑇𝑏 approaches zero, the time 7 \n evolution of 𝒎̂ diverg es if 𝐽𝑠 remains finite at xcJ. As observed in Fig. 4(a), 𝜒𝑠𝑎𝑡 remains roughly \nunchanged across xcJ, suggesting that similar to the field -driven experiment, SOT could also be used as an \nefficient drive force for achieving fast dynamics at this concentration. Finally, we find the switching \npolarity keeps the same sign across xcJ, differ ing from the current induced switching of CoGd spin valves \nstudied in Ref. [29], where a sw itching polarity reversal was observed between xcJ and xcM. This \ndifference is due to the presence of different switching mechanisms: in SOT induced switching of PMA \nfilms, the two competing torques are the field torque 𝛾𝑒𝑓𝑓𝑴×𝑯𝑒𝑓𝑓 and the spin tor que. Because the two \nterms have the same pre -factor 𝛾𝑒𝑓𝑓 and are only functions of 𝑴, 𝑯𝑒𝑓𝑓, and 𝝈̂, under the same applied 𝝈̂ \nand 𝑯𝑦, the orientation of 𝒎̂ will remain the same (Fig. 2b ), regardless of the sign of 𝛾𝑒𝑓𝑓. In contrast, \nthe anti -damping switching of spin valves changes polarity for regions with 𝛾𝑒𝑓𝑓<0, as explained in \nRef. [29]. \n \nV. Measurement of the Dzyaloshinskii -Moriya I nteraction Energy \n The in -plane field needed for saturating the SOT, 𝐻𝑦𝑠𝑎𝑡, is plotted against the Tb concentration in \nFig. 4(b). First of all, we notice that 𝐻𝑦𝑠𝑎𝑡 is largest near xcM. This result is consistent with fact that the \neffective DMI field [32] 𝐻𝐷𝑀𝐼 =𝐷/𝑀𝑠𝑡𝜇0∆, where 𝐷 is the DMI energy density and ∆ is the DW width, \nwould become divergent when 𝑀𝑠 approaches zero . Secondly, 𝐻𝑦𝑠𝑎𝑡 is generally larger for the Tb-\ndominant sample s than the Co -dominant ones. For example, the sample with the highest Co \nconcentration, Co0.87Tb0.13, shows 𝐻𝑦𝑠𝑎𝑡~100 Oe, which is close to the reported saturation field of Ta/FM \nstacks [25]. However, in the Tb dominant sample Co0.71Tb0.29, which has similar magnetic moment, 𝐻𝑦𝑠𝑎𝑡 \nis found to be ~1500 Oe. In the inset of Fig. 4(b) we plot 𝐻𝑦𝑠𝑎𝑡𝑀𝑠𝑡, which increases roughly linearly as a \nfunction of x. By calculating the DW width ∆ =√𝐴𝐾𝑢⁄ using the determined anisotropy energy 𝐾𝑢 = 6.4 \n× 104 J∙m-2 from Tb - and Co -dominant samples and the reported exchange stiffness [39] A ~ 1.4 × 10-11 8 \n J/m, we get D in the range of 0.05~0.66 pJ/m. The increasing DMI energy with increasing Tb \nconcentration can be explained by the strong spin-orbit coupling and large deviation from the free \nelectron g factor [30] in the Tb atoms. The generation of magnetic textures such as chiral DWs and \nmagnetic skyrmion s [40] relies on the competition between the DMI energy and other magnetostatic \nenergies. Therefore, the tunable DMI through chemical composition provides a useful a knob for \ncontrolling magnetic phases. \n \nVI. Conclusion \n To summarize, we demonstrated SOT induced switching in Co1-xTbx thin films with a wide range \nof chemical compositions. The effective field from the SOT was found to scale with the inverse of \nmagnetic moment, consistent with the conservation of angular mome ntum. The high efficiency of SOT at \nthe compensation points as well as the previously demonstrated fast dynamics in these systems makes \nthem highly attractive for high speed spintronic applications . Moreover , we found that the DMI energy \ndensity is much la rger in samples with high rare earth concentrations, which could provide useful \napplications in spintronic devices that employ stable magnetic textures. 9 \n References \n[1] T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich, “Antiferromagnetic spintronics,” Nat. \nNanotechnol. 11, 231 (2016). \n[2] C. Marrows, “Addressing an ant iferromagnetic memory,” Science 351, 558 (2016). \n[3] I. A. Campbell,“Indirect exchange for rare earths in metals”, J. Phys. F Met. Phys. 2, L47 (1972). \n[4] K. Lee and N. Heiman,“Magnetism in rare earth‐transition metal amorphous alloy films,” in AIP \nConf. Proc. (AIP Publishing, 1975), pp. 108 –109. \n[5] K. H. J. Buschow,“Magnetic properties of amorphous rare‐earth –cobalt alloys,” J. Appl. Phys. 51, \n2795 (1980). \n[6] Y. J. Choe, S. Tsunashima, T. Katayama, and S. Uchiyama, “Magneto -optic Kerr spectra of \namorpho us RE -Co thin films,” J. Magn. Soc. Jpn. 11, S1_273 (1987). \n[7] P. Hansen, C. Clausen, G. Much, M. Rosenkranz, and K. Witter, “Magnetic and magneto‐optical \nproperties of rare‐earth transition‐metal alloys containing Gd, Tb, Fe, Co,” J. Appl. Phys. 66, 756 \n(1989). \n[8] C. D. Stanciu, A. V. Kimel, F. Hansteen, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing,“Ultrafast \nspin dynamics across compensation points in ferrimagnetic GdFeCo: The role of angular \nmomentum compensation,” Phys. Rev. B 73, 220402 (2006). \n[9] M. Binder, A. Weber, O. Mosendz, G. Woltersdorf, M. Izquierdo, I. Neudecker, J. R. Dahn, T. D. \nHatchard, J. -U. Thiele, C. H. Back, and M. R. Scheinfein, “Magnetization dynamics of the \nferrimagnet CoGd near the compensation of magnetization and angular momentum,” Phys. Rev. B \n74, 134404 (2006). \n[10] M. I. Dyakonov and V. I. Perel, “Current -induced spin orientation of electrons in semiconductors,” \nPhys. Lett. A 35, 459 (1971). \n[11] J. E. Hirsch, “Spin Hall Effect,” Phys. Rev. Lett. 83, 1834 (1999). \n[12] Y. A. Bychkov and E. I. Rashba,“Properties of a 2D electron gas with lifted spectral degeneracy,” \nJETP Lett 39, 78 (1984). \n[13] V. M. Edelstein,“Spin polarization of conduction electrons induced by electric current in two -\ndimensional asymmetric electron systems,” Solid State Commun. 73, 233 (1990). \n[14] I. M. Miron, K. Garello, G. Gaudin, P. -J. Zermatten, M. V. Costache, S. Auffret , S. Bandiera, B. \nRodmacq, A. Schuhl, and P. Gambardella, “Perpendicular switching of a single ferromagnetic layer \ninduced by in -plane current injection,” Nature 476, 189 (2011). \n[15] L. Liu, C. -F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman,“S pin-Torque Switching with the \nGiant Spin Hall Effect of Tantalum,” Science 336, 555 (2012). \n[16] Y. Fan, P. Upadhyaya, X. Kou, M. Lang, S. Takei, Z. Wang, J. Tang, L. He, L. -T. Chang, M. Montazeri, \nG. Yu, W. Jiang, T. Nie, R. N. Schwartz, Y. Tserkovnyak, a nd K. L. Wang,“Magnetization switching \nthrough giant spin –orbit torque in a magnetically doped topological insulator heterostructure,” \nNat. Mater. 13, 699 (2014). \n[17] H. Kurebayashi, J. Sinova, D. Fang, A. C. Irvine, T. D. Skinner, J. Wunderlich, V. Novák , R. P. \nCampion, B. L. Gallagher, E. K. Vehstedt, L. P. Zârbo, K. Výborný, A. J. Ferguson, and T. \nJungwirth,“An antidamping spin -orbit torque originating from the Berry curvature,” Nat. \nNanotechnol. 9, 211 (2014). \n[18] Z. Zhao, M. Jamali, A. K. Smith, and J.-P. Wang,“Spin Hall switching of the magnetization in \nTa/TbFeCo structures with bulk perpendicular anisotropy,” Appl. Phys. Lett. 106, 132404 (2015). \n[19] Y. Mimura, N. Imamura, and Y. Kushiro,“Hall effect in rare‐earth –transition‐metal amorphous alloy \nfilms,” J. Appl. Phys. 47, 3371 (1976). \n[20] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong,“Anomalous Hall effect,” Rev. \nMod. Phys. 82, 1539 (2010). 10 \n [21] P. P. J. Haazen, E. Murè, J. H. Franken, R. Lavrijsen, H. J. M. Swagten, and B. Koopm ans, “Domain \nwall depinning governed by the spin Hall effect,” Nat. Mater. 12, 299 (2013). \n[22] S. Emori, U. Bauer, S. -M. Ahn, E. Martinez, and G. S. D. Beach,“Current -driven dynamics of chiral \nferromagnetic domain walls,” Nat. Mater. 12, 611 (2013). \n[23] S. Emori, E. Martinez, K. -J. Lee, H. -W. Lee, U. Bauer, S. -M. Ahn, P. Agrawal, D. C. Bono, and G. S. D. \nBeach,“Spin Hall torque magnetometry of Dzyaloshinskii domain walls,” Phys. Rev. B 90, 184427 \n(2014). \n[24] O. J. Lee, L. Q. Liu, C. F. Pai, Y. Li, H. W. Tseng, P. G. Gowtham, J. P. Park, D. C. Ralph, and R. A. \nBuhrman,“Central role of domain wall depinning for perpendicular magnetization switching driven \nby spin torque from the spin Hall effect,” Phys. Rev. B 89, 24418 (2014). \n[25] C.-F. Pai, M. Mann, A. J . Tan, and G. S. D. Beach, “Determination of spin torque efficiencies in \nheterostructures with perpendicular magnetic anisotropy,” Phys. Rev. B 93, 144409 (2016). \n[26] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and R. A. Buhrman, “Current -Induced Sw itching of \nPerpendicularly Magnetized Magnetic Layers Using Spin Torque from the Spin Hall Effect,” Phys. \nRev. Lett. 109, 96602 (2012). \n[27] J. C. Slonczewski, “Current -driven excitation of magnetic multilayers,” J. Magn. Magn. Mater. 159, \nL1 (1996). \n[28] H. V. Gomonay and V. M. Loktev, “Spin transfer and current -induced switching in \nantiferromagnets,” Phys. Rev. B 81, 144427 (2010). \n[29] X. Jiang, L. Gao, J. Z. Sun, and S. S. P. Parkin, “Temperature Dependence of Current -Induced \nMagnetization Switching in Spin Valves with a Ferrimagnetic CoGd Free Layer,” Phys. Rev. Lett. 97, \n217202 (2006). \n[30] T. Moriya, “Anisotropic Superexchange Interaction and Weak Ferromagnetism,” Phys. Rev. 120, 91 \n(1960). \n[31] T. Tono, T. Taniguchi, K. -J. Kim, T. Moriyama, A. Tsukam oto, and T. Ono, “Chiral magnetic domain \nwall in ferrimagnetic GdFeCo wires,” Appl. Phys. Express 8, 73001 (2015). \n[32] A. Thiaville, S. Rohart, É. Jué, V. Cros, and A. Fert, “Dynamics of Dzyaloshinskii domain walls in \nultrathin magnetic films,” EPL Europh ys. Lett. 100, 57002 (2012). \n[33] J. Z. Sun and D. C. Ralph,“Magnetoresistance and spin -transfer torque in magnetic tunnel \njunctions,” J. Magn. Magn. Mater. 320, 1227 (2008). \n[34] P. M. Haney, R. A. Duine, A. S. Núñez, and A. H. MacDonald,“Current -induced torques in magnetic \nmetals: Beyond spin -transfer,” J. Magn. Magn. Mater. 320, 1300 (2008). \n[35] C.-F. Pai, Y. Ou, L. H. Vilela -Leão, D. C. Ralph, and R. A. Buhrman,“Dependence of the efficiency of \nspin Hall torque on the transparency of Pt/ferromagnetic la yer interfaces,” Phys. Rev. B 92, 64426 \n(2015). \n[36] B. I. Min and Y. -R. Jang, “The effect of the spin -orbit interaction on the electronic structure of \nmagnetic materials,” J. Phys. Condens. Matter 3, 5131 (1991). \n[37] J. M. D. Coey, Rare -Earth Iron Perman ent Magnets (Clarendon Press, 1996). \n[38] R. K. Wangsness,“Sublattice Effects in Magnetic Resonance,” Phys. Rev. 91, 1085 (1953). \n[39] J. J. Turner, X. Huang, O. Krupin, K. A. Seu, D. Parks, S. Kevan, E. Lima, K. Kisslinger, I. McNulty, R. \nGambino, S. Mang in, S. Roy, and P. Fischer, “X -Ray Diffraction Microscopy of Magnetic Structures,” \nPhys. Rev. Lett. 107, 33904 (2011). \n[40] X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Matsui, N. Nagaosa, and Y. Tokura,“Real -\nspace observation of a two -dimens ional skyrmion crystal,” Nature 465, 901 (2010). \n \n 11 \n \n \nFig. 1 (a) Out of plane magnetization curves of Co 1-xTbx films . (b) RAH as a function of perpendicular \nmagnetic field. (c ) Magnetic momen ts of Co1-xTbx alloys as a function of Tb concentration . (d) Schematic \nof the device geometry for RAH measurement. (e) RAH as a function of Tb concentration . \n \n \n \n \n \n \n \n \n \n \n12 \n \n \nFig. 2 (a) Current induced SOT switching of Co 1-xTbx for in -plane fields of ± 2000 Oe. The current \ndensity inside Ta is calculated based on the conductivity of Ta thin films. (b) Schematic of effective fields \nin a ferrimagnetic system. Fields acting on the moment consist of the in -plane field Hy, the anisotropy \nfield Han, and the SOT field HST. \n \n \n13 \n \n \nFig. 3 (a), (c),(e) Measurements on Co-dominant sample Co 0.82Tb0.18. (b),(d), (f) Measurements on Tb-\ndominant sample Co 0.75Tb0.25. (a),(b ) RAH vs. applied perpendicular field under a DC current of ±3 mA. \n(c),(d ) SOT effective field as a function of applied current density under in -plane fields of ±2000, and 0 \nOe. (e ),(f) SOT efficiency vs. Hy. Efficiency saturates at the field Hsaty. (g) SOT induced DW motion in \nthe ferrimagnetic system for Tb-dominant and Co-dominant films , showing that the effective \nperpendicular field has the same sign in both cases. \n14 \n \n \nFig. 4 (a ) Saturation efficiency and (b) in-plane saturation field for differe nt Co 1-xTbx films. Both the \nsaturation efficiency and in-plane saturation field are largest near the magnetic moment compensation \npoint. The dashed line in (a) shows the trend calculated from χsat ~1/ Mst. Inset of (b) plots the product of \nthe in -plane sat uration field and magnetic moment M st, indicating an increase in DMI energy density D \nwith increasing Tb concentration. \n \n \n \n" }, { "title": "2009.10222v1.Magneto_Elastic_Coupling_to_Coherent_Acoustic_Phonon_Modes_in_Ferrimagnetic_Insulator_GdTiO__3_.pdf", "content": "arXiv:2009.10222v1 [cond-mat.str-el] 21 Sep 2020Magneto-Elastic Coupling to Coherent Acoustic Phonon\nModes in Ferrimagnetic Insulator GdTiO 3\nD. Lovinger,1E. Zoghlin,2P. Kissin,1G. Ahn,3K. Ahadi,2P. Kim,1M.\nPoore,1S. Stemmer,2S. J. Moon,3S. D. Wilson,2and R. D. Averitt1\n1Department of Physics, University of California, San Diego , La Jolla, CA, 92093\n2Materials Department, University of California, Santa Bar bara, CA, 93106\n3Department of Physics, Hanyang University, Seoul, South Ko rea, 04763\n(Dated: September 23, 2020)\nIn this work we investigate single crystal GdTiO 3, a promising candidate material for Floquet en-\ngineering and magnetic control, using ultrafast optical pu mp-probe reflectivity and magneto-optical\nKerr spectroscopy. GdTiO 3is a Mott-Hubbard insulator with a ferrimagnetic and orbita lly ordered\nground state ( TC= 32 K). We observe multiple signatures of the magnetic phase transition in\nthe photoinduced reflectivity signal, in response to above b and-gap 660 nm excitation. Magnetic\ndynamics measured via Kerr spectroscopy reveal optical per turbation of the ferrimagnetic order\non spin-lattice coupling timescales, highlighting the com petition between the Gd3+and Ti3+mag-\nnetic sub-lattices. Furthermore, a strong coherent oscill ation is present in the reflection and Kerr\ndynamics, attributable to an acoustic strain wave launched by the pump pulse. The amplitude of\nthis acoustic mode is highly dependent on the magnetic order of the system, growing sharply in\nmagnitude at TC, indicative of strong magneto-elastic coupling. The drivi ng mechanism, involving\nstrain-induced modification of the magnetic exchange inter action, implies an indirect method of\ncoupling light to the magnetic degrees of freedom and emphas izes the potential of GdTiO 3as a\ntunable quantum material.\nI. INTRODUCTION\nTherare-earthtitanates (unit formulaRTiO 3, whereR\nis a rare-earth ion) are a class of complex materials with\nstrongly correlated spin, orbital, and lattice degrees of\nfreedom. They are3 d1compounds with a single d-orbital\nelectron occupying the Ti3+t2gorbital, whose degener-\nacy is broken by strong crystal field splitting [1]. This\npresents an opportunity to study a strongly correlated\nsystem in relative simplicity, which nonetheless exhibits\nrich physics and interesting properties. The perovskiteti-\ntanates, for example, are Mott-Hubbard (MH) insulators\nwith interconnected orbital and spin order [2–6]. Of par-\nticular interest is the complex magnetic phase diagram\nfor this class of materials, with a magnetic ground state\nthat varies from ferrimagnetic to antiferromagnetic as a\nfunction of the rare-earthion size and subsequent change\nin Ti-O-Ti bond angle [1, 7]. Various theories have at-\ntempted to explain the magnetic order in titanates [5, 8–\n10], all of which highlight the need to consider the roles\nof structure and electronic correlation to understand the\ncomplexity embodied in the magnetic phase diagram.\nCommon to all descriptions of magnetism in the per-\novskite titanates is the defining role of the lattice and\nits distortion. It has been argued, for example, that the\ndegree of GdFeO 3distortion and changes to the Ti-O-\nTi bond angle directly modify the exchange interaction\nwhich, in turn, determines the magnetic order [7, 11].\nMore recent results emphasize the importance of orbital\norder in determining the magnetic order. In particular,\nthe direct coupling between the orbital order and lat-\ntice, ratherthantheorthorhombicdistortion, contributes\nmost strongly to the ground state [12–14]. Whether itis particular structural distortions or more generalized\nJahn-Teller distortions, and regardless of the role of or-\nbital ordering, it is clear that magnetic order in titanates\nis highly dependent on the lattice.\nIn this work we study GdTiO 3(GTO), a titanate with\nan orthorhombic perovskite-type unit cell and relatively\nlarge GdFeO 3-type distortion. GTO lies just within the\nferromagnetic (FM) region of the phase diagram. The\nproximity to the FM-AFM transition makes GTO partic-\nularly sensitive to the effect of structural changes on the\nmagnetism [12]. Below the critical temperature TC=\n32 K it is ferrimagnetically (fM) ordered: the Ti3+spins\nare aligned ferromagnetically along the c-axis and cou-\npled antiferromagnetically to the Gd sublattice [3, 8, 12].\nThe magnetism saturates at 6 µB(7µBGd – 1µBTi) in\na relatively small field of ∼0.1 T, with no discernable hys-\nteresis [15]. The magnetocrystalline anisotropy is small,\nwith the a-axis as the hard magnetization axis and the\nb-cplane nearly isotropic. The fM order is accompanied\nand mediated by ( yz, zx, yz, zx )-type orbital order, a re-\nsult of inter-atomic hybridization between the t 2gand eg\norbitals [1].\nThepresentworkonGTOis motivated notonlyby the\nrelative simplicity of the system and rich interconnected\norder, but also the potential for Floquet engineering and\nultrafast control of magnetism. Liu et al.explored the\nMott insulating titanates as a candidate for tuning the\nspin-orbital Floquet Hamiltonian and subsequent modi-\nfication of the spin exchange interaction using light [16].\nMeanwhile, Khalsa et al.suggest direct excitation of a\nGTO mid-IR active phonon mode to transiently mod-\nify the exchange interaction and switch the ground state\nfrom FM to AFM on ultrafast timescales [17]. A similar2\n0 2 4 6 \nPhoton Energy (eV )0123451()( -1 cm -1 )\n10 \n50 \n150\n3001 2 3 \nPhoton Energy (eV) 120160200240\n0 2 4 6 \nPhoton Energy (eV )0123456n( )1 2 3 \nPhoton Energy (eV) 2.12.22.310 3(a) (b)\nTi 1 t2g LHBTi 2 t2g UHB\nO 2pEFU\n1.88 eV(c)\nT (K) \n10 \n50 \n150\n300T (K) 1.88 eV\nFIG. 1. (a) Optical conductivity and (b) index of refraction of GTO/LSAT thin film as a function of photon energy and\ntemperature. Red arrows indicate the 1.88 eV pump/probe ene rgy. The weak feature in σ1at 2 eV corresponds to the MH\ngap, while the steep feature near 5 eV arises from O 2pto Ti3dand Gd 4fcharge transfer transitions. (c) Depiction of 1.88 eV\nlaser excitation, corresponding to intersite Ti 3d-3dtransition across the MH gap.\nexperiment utilizing phononic control has been proposed\nby Guet al.in other titanates [18].\nWhile the conditions of our experiment lie outside\nthe regimes discussed above, we do observe strong cou-\npling between light, the lattice, and the sample mag-\nnetism. Time-resolved pump-probe and magneto-optical\nKerr effect (MOKE) measurements tuned to ∼1.88 eV,\njust above the bandgap, allow us to measure the evo-\nlution of photoexcited states on femtosecond – picosec-\nond timescales. We observe multiple signatures of the\nmagnetic phase transition in the photoinduced reflectiv-\nity signal, as well as optical perturbation of the fM order\non spin-lattice coupling timescales in the MOKE signal.\nIn addition, an acoustic phonon mode is present in both\nsignals, whose amplitude is highly coupled to the mag-\nnetic order. This implies strong magneto-elasticcoupling\nthrough transient, strain-induced modification of the ex-\nchange interaction, connecting the lattice and magnetic\ndegrees of freedom and indicating that the exchange in-\nteraction is tunable on ultrafast timescales.\nII. METHODS\nSingle crystaland thin film samplesofGdTiO 3werein-\nvestigated. The photoinduced reflectivity signal in both\nis extremely similar and the following work, except for\nthe measurement of the optical constants, was performed\non a single crystal sample. For comparison, the thin\nfilm time-resolved reflectivity data is presented in SM\nI [19]. GdTiO 3thin films ( ∼20 nm) were grown on\na (001)(La 0.3Sr0.7)(Al0.65Ta0.35)O3(LSAT) substrate by\nhybrid molecular beam epitaxy [20]. GdTiO 3bulk sin-\ngle crystals were grown by high pressure laser floating\nzone method [21]. A small fraction of the crystal rodwas cut and polished to optical quality, with bc-axis in\nplane and a-axis out of plane. Powder X-ray diffraction\nmeasurements indicate extremely high quality crystals\nwith no notable impurity peaks and lattice parameters\nat 5.393, 5.691, 7.664 ˚A fora,b,c-axis [21], well matched\nto literature values [12]. Magnetization measurements in-\ndicate no visible hysteresis and a saturation moment of\n6µB/FU.\nTo determine the optical conductivity and index of re-\nfraction, frequency-dependent reflectivity spectra R(ω)\nin the photon energy region between 3 meV and 85\nmeV were measured by using a Bruker VERTEX 70v\nFourier transform spectrometer. The GdTiO 3thin film\nwas mounted in a continuous liquid helium flow cryostat.\nWe used two spectroscopic ellipsometers (IR-VASE Mark\nII and M-2000, J. A. Woollam Co.) for obtaining the\ncomplex dielectric constants ǫ(ω) =ǫ1(ω) +iǫ2(ω) in\nthe energy range from 60 meV to 0.75 eV and 0.75 eV\nto 6.4 eV, respectively. The optical conductivity of the\nGdTiO 3film was obtained by two-layermodel fit employ-\ning Drude-Lorentz oscillators for optical response of each\nlayer [22].\nUltrafast optical pump-probe reflectivity measure-\nments (∆R/R) are performed using a 1040 nm 200 kHz\nSpectra-Physics Spirit Yb-based hybrid-fiber laser cou-\npled to a non-colinear optical parametric amplifier. The\namplifierproduces ∼20fs pulsescenteredat660nm (1.88\neV), which are split, cross-polarized (pump s-polarized,\nprobep), and used asdegeneratepump and probebeams.\nThe pump is aligned along the b-axis of the GTO crystal.\nThis excitation corresponds to an intersite Ti 3d–3d\ntransition across the Mott-Hubbard gap, shown in Fig.\n1(c). A moderate pump fluence of ∼100µJ/cm2is used\nto minimize sample heating ( ∼4 K at 10 K), ensuring we\nare in the linear excitation regime.3\nTime-resolved magneto-optical Kerr spectroscopy is\nused to probe the magnetization dynamics. The same op-\ntical system described above is used here, including laser\nenergy, fluence, and magneto-optical cryostat (Quan-\ntum Design OptiCool). The photoinduced Kerr rotation\n(∆θK) is measured using balanced photodiodes in the po-\nlar Kerr geometry at near-normalincidence, in a continu-\nously variable external magnetic field (0 – 7 T), with the\npump polarized along the b-axis and Kerr probe polar-\nized along the c-axis of the crystal. The magnetic field is\napplied normal to the sample surface, along the a-axis of\nthe crystal, resulting in a Kerr signal proportional to the\nout-of-plane z-component of the photoinduced change in\nmagnetization ∆ Mz. In order to eliminate non-magnetic\ncontributions to the signal and ensure we are measur-\ning genuine spin dynamics, we take the difference of the\nKerrsignalatvariouspositive andnegativeapplied fields:\n∆θK= ∆θ(+M)−∆θ(−M) (see SM IV for details)\n[19, 23–27].\nIII. EXPERIMENTAL RESULTS\nThe temperature-dependent optical conductivity of\nthin-film GdTiO 3is shown in Fig. 1(a) for photon ener-\ngies ranging from 3 meV to 6.5 eV, and for temperatures\nfrom 10 K to 300 K. A weak feature is present, centered\nat 2 eV, corresponding to the Mott-Hubbard gap. Apart\nfrom weak thermal broadening with increasing temper-\nature, the peak at 2 eV is nearly temperature indepen-\ndent. While early studies of GTO measured a MH gap\nof 0.2 – 0.7 eV [28], more recent photoluminescence and\nDFT/DFT+U results place the gap closer to 1.8 – 2 eV\n[29]. Thesmallpeakintheopticalconductivityspectrum\nat 2 eV measured here supports these recent findings. At\nmuch higher energies we observe a significant increase in\nthe optical conductivity. The features near 5 eV corre-\nspond to O 2pto Ti3dand Gd 4ftransitions. Fig. 1(b)\nshows the index of refraction in the same energy range,\nwhich remains relatively constant as a function of tem-\nperature.\nThe time-dependent photoinduced change in reflectiv-\nity ∆R/Rfor a GdTiO 3single crystal is shown in Fig.\n2(a), for all measured temperatures between 10 – 295\nK (legend on Fig. 2(b)). The black lines represent ex-\nponential fits to the data as described below. The pho-\ntoinduced change in ∆ R/Ris positive; following laser\nexcitation a non-equilibrium electron population is estab-\nlished in ∼500 fs, which then exchanges energy and equi-\nlibrates with the spin and lattice subsystems throughvar-\nious pathways, each with a characteristic timescale. This\nis visible as the slower, multi-component exponential re-\nlaxation. As the temperature is decreased from 295 K\nthe signal amplitude increases, recovery dynamics slow,\nandtwoadditionalfeaturesemerge. The firstisa delayed\nrise time, correspondingto a further departure from equi-\nlibrium in the first ∼15 ps, emerging below T= 100 K.0 200 400 600 800 1000 \nTime (ps) 010 20 R/R 10 -3 \n0 5 10 15 \nTime (ps) 0510 R/R 10 -3 0 500 1000\nTime (ps) 0510 15 20 R/R 10 -3 \n10 K\nslow e-ph \ns-l OO \n100 K0 100 200 300\nTemperature (K) 0510 15 20 25 Peak R/R (500 ps) 10 -3 \n295\n200\n100\n80 \n60 \n45 \n40 \n35 \n30 \n25 \n20 \n10 (a) (b)\n(c)T (K) \nFIG.2. (a)Photoinduceddifferentialreflectivitysignal∆ R/R\nat all measured temperatures, from 10 K to 295 K, taken on\na single crystal GdTiO 3sample. The black curves are expo-\nnential fits to the data, of the form given in Eq. 1. The\nblue star indicates the data curve taken at TC. (b) ∆R/R\nvalues at 500 ps, an approximation of the peak signal at all\ntemperatures. The red line is a power law fit, commonly seen\nin systems undergoing a second-order magnetic phase transi -\ntion. (c) Representative pump-probe scans at 10 K and 100\nK, indicating the various timescales involved in the recove ry\nprocess (see Eq. 1).\nSecond, there is a crossover point visible at delay times\nof∼200 ps where recovery dynamics flatten and reverse\ndirection to become an additional rise time. This occurs\nprecisely as the ferrimagnetic ordering temperature TC\n= 32 K is crossed (marked by a blue star), indicating\nthat magnetization dynamics manifest in the differential\nreflectivity signal.\nTo further investigate the temperature dependence of\nthe reflectivity signal, we plot the peak signal amplitude\n(at 500 ps) in Fig 2(b). The behavior here is distinctive,\nnot uncommon in materials undergoing a second-order\nmagnetic phase transition. The red curve represents a\npower-law fit to the data, of the form A=A0t−w, where\nwis the critical exponent and tis the reduced tempera-\nturet=|T−TC|\nTC. The value of wdepends upon the sym-\nmetry and universality class of the magnetic transition.\nOurfit producesacriticalexponent w= 1.28±0.02. This\nvery nearly matches the critical behavior predicted by\ndynamical scaling theory for the 3D Ising model, which\nyieldsw≈1.32[30–32](furtherdetailedinSMII)[19,30–\n35]. Whilethisisanindirectmethodofmeasuringcritical\ndynamics and is not intended to be a rigorous analysis,4\n050 100 150 s-l (ps) \n-6 -4 -2 0\nAmplitude s-l 10 -3\n 50\n 60\n 80\n100\n150\n250 10\n 20\n 25\n 30\n 35\n 400 100 200 300 \nTemperature (K) \n0\nTime (ps) -3 -2 -1 01R/R Residual 10 -4 \n0 50 100 150\nTime (ps) -2 -1 01R/R Res. 10 -4 (a)\n(b)\n10 20 30 40 50 T (K) \nFIG. 3. (a) Time constant (black) and amplitude (red) of the\nspin-lattice coupling term, extracted from exponential fit s to\nthe ∆R/Rdata. The vertical gray section indicates the fM\ntransition region TC= 32 K, where the lifetime is too long\nto measure (see 30 K curve marked by a blue star in Fig.\n2(a)). The reddashed line depicts zero amplitudeand clarifi es\nthe crossover region. (b) Coherent acoustic phonon respons e,\nisolated by subtracting the exponential fits from the ∆ R/R\ndata. The inset shows the ∆ R/Rresidual to a longer delay\ntime of 150 ps, where higher frequency components emerge.\nit is clear that the peak reflectivity follows power-law be-\nhavior as expected at a magnetic phase transition. The\nresults suggests that there is indeed a magnetic contribu-\ntion in the ∆ R/Rsignal. Additionally, the qualitative\nform of the peak amplitude vs temperature follows that\nof the temperature dependent magnetic susceptibility in\nbulk GTO [15], and the magnetization M in films [21, 36].\nWhile by no means conclusive, the universal scaling be-\nhavior and agreement with thermal magnetization does\nstrongly suggest that the ∆ R/Rsignal measured, par-\nticularly at longer times (500+ ps), is sensitive to spin\ndynamics.\nTosubstantiatetheseclaims, wequantitativelyanalyze\nthe full time-dependent response. Below 100 K, the dy-\nnamics can be fit by a sum of four exponentials with a\nconstant offset, of the form:∆R/R(t) =Ae−phe−t/τe−ph+AOOe−t/τOO+\nAs−le−t/τs−l+Aslowe−t/τslow+C,(1)\nshown as black lines in Fig. 2(a). Not listed is an ad-\nditional error function term, which describes the initial\nstep-like rise dynamics at t= 0. A visual representation\nofthevarioustimescalesisshowninFig. 2(c)fortwotem-\nperatures. After excitation the dynamics follow a general\ntrend; there is a very fast initial recovery, τe−phon the\norder of ∼500 fs, followed by an intermediate term τOO\non the order of 2 – 8 ps, both of which are clearly visible\nin the inset of Fig. 2(c). Note that τOOis an additional\nrise time which vanishes at higher temperatures, the full\ndynamics fitting to only 3 exponentials (i.e. above TC).\nThis is followed by a slower term τs−lon the order of\n100’s of picoseconds, and a final much slower recovery\nτslow. A careful inspection of the reflectivity data also in-\ndicates the presence of small oscillations about the black\nfitted curves, which we discuss below.\nThese measured timescales are well separated and can\nbe attributed to distinct physical processes. The initial\npump pulse excites an intersite Ti 3d-3dtransition. This\ndirectly creates a population of hot carriers which ther-\nmalize via electron-electron (e-e) scattering, then subse-\nquently exchange energy with the lattice, orbital, and\nspin degreesof freedom. We focus on the spin-lattice cou-\npling process here, with a full discussion of the remaining\nprocesses and time constants in SM III [1, 11, 19, 37–45].\nThe most relevant component of the ∆ R/Rsignal\nis the third fitted exponential, τs−l, attributed to spin-\nlattice coupling and shown in Fig. 3(a). This term has a\ncharacteristic timescale of 10 – 140 ps, excluding the re-\ngion at the magnetic phase transition temperature TC=\n32K where the lifetime growstoo long to accuratelymea-\nsure. This critical region is visible as a flattening of the\n∆R/R recoveryat 30 K, indicated by the blue star in Fig.\n2(a). Before the onset of this third recovery term τs−l,\nthe ∆R/R signal reveals dynamics indicative of electron-\nlattice equilibration. It follows that this longer lifetime\nis related to equilibration of the spin subsystem with the\nlattice. The time constant measured, on the order of 100\nps in the magnetic phase, is consistent with spin-lattice\ncoupling in other magnetic insulators [46–48]. The char-\nacteristic time is relatively constant in the paramagnetic\nphase until 150 −200 K, where it begins to slowly in-\ncrease. This corresponds to the onset temperature ( ∼180\nK) of spin-spin coupling between the Gd3+and Ti3+ions\n[8]. Closer to 100 K τs−lfurther increases, indicating the\nonset of short-range fM spin correlations. This is also\napparent in the increase in amplitude at this tempera-\nture. Finally, as TCis crossed (dark gray region) we see\nevidence of the second-order ferrimagnetic phase transi-\ntionasthetimeconstantdivergesandamplitudeswitches\nsign. The now-negative amplitude implies an additional\nrise time in the signal; as energy is transferred to spins\nand the ferrimagnetic order is disrupted, the system is5\n-2 0246 (rad) 10 -4 \n-2 0246\n0 500 1000 \nTime (ps)-2 0246 (rad) 10 -4 \n0 500 1000\nTime (ps) 0246810 10 -4 \n 10\n 20\n 25\n 30\n 35\n 40\n 50\n 60\n 80\n25010 -4 H = 0.1 T 0.25 T\n0.5 T1 TT (K) \nFIG. 4. Time-resolved Kerr dynamics at various magnetic\nfields, recorded as the difference between the MOKE signal\nin opposing field directions ∆ θ= ∆θ(+H)−∆θ(−H). This\nis a measure of the photoinduced change in the out-of-plane\nmagnetization ∆ Mz. Red arrows indicate the crossover to\nnegative values of ∆ θ.\nbrought further out of equilibrium. In the paramagnetic\nphase there is no long-range spin order to disrupt, such\nthat spin-lattice thermalization manifests as a simple re-\ncovery to equilibrium. The critical behavior, amplitude\nreversal, timescale, and temperature dependence of the\nτs−lcomponent all suggest that we are measuring spin-\nlattice coupling on a timescale of ∼100 ps, and that it is\nhighly sensitive to the onset of magnetic order.\nThe final interesting feature of the ∆ R/Rdata is a\nslow coherent oscillation, prominent at early times. By\nsubtracting the exponential fits at each temperature we\ncan extract the oscillatory component, plotted in Fig.\n3(b). The result is peculiar – we observe a low-frequency\nphonon mode which grows in amplitude and becomes\nchirped, slowing down and redshifting as it propagates.\nThe oscillationperiod (on the orderof20ps), suggestsan\nacoustic strain wave launched by the pump pulse which\npropagates through the crystal [49]. The probe beam re-\nflected from the sample surface interferes with a portion\nreflected from the strain wave boundary, resulting in an\noscillatory signal. The temperature dependence of this\nmode is striking – the amplitude is relatively constant at\nhigh temperatures, then grows sharply precisely at the\nfM phase transition temperature. Though it appears to\nbe an acoustic mode, it is also clearly coupled to the\nmagnetic order. This suggests strong magneto-acoustic\ncoupling, tying the dynamics of the magnetic subsystem\nto the transiently strained lattice.\nTo gain further insight into the magnetization dynam-\nics and the influence upon acoustic phonon propaga-tion, we utilize time-resolved magneto-optical Kerr effect\n(MOKE) spectroscopy. Fig. 4presents the photoinduced\nKerr rotation ∆ θ, proportional to the change in out-of-\nplane (a-axis) magnetization ∆ Mz, for all temperatures\nand four fields between 0.1 – 1 T. For details of the anal-\nysis, see SM IV [19, 23–27]. Additional static Kerr rota-\ntion measurements are presented in SM V [19]. At lower\nfield strengths, the Kerr signal reveals a quick rise in\nthe photoinduced out-of-plane magnetization ∆ Mz, fol-\nlowed by a reduction and change in sign of ∆ Mz. This\ncan be interpreted asa pump-induced increaseand subse-\nquent decrease in the net out-of-plane magnetic moment,\nbut not necessarily a reversal of the total magnetic mo-\nment. Therearetwoprimarycomponentsto the Kerrsig-\nnal, one positive (growing in ∼100 ps), and one negative\n(growing in slower, ∼100 – 500 ps). These dynamics are\nslow and long-lived, as expected in magnetic insulators\nlikeGTOduetothelocalizednatureofquasiparticles[48].\nTo describethe temperature dependence ofthe signal, we\nfocus on lower field strengths H= 0.1−0.5 T. At high\ntemperature, in the paramagnetic phase, the Kerr signal\nis weak and indicates the lack of long-range magnetic or-\nder. As the temperature is lowered there is an increase in\nthe photoinduced rotation, with a clear negative signal\nemerging below TC. This negative component is largest\nand appears at earlier delays right at the transition tem-\nperature ( TC= 32 K). With decreasingtemperature, the\ncrossover to negative values of ∆θoccurs at later times.\nWell below TC, in the strongly ordered phase, the signal\nremains positive at all time delays.\nWe also observe a significant field dependence in the\ndata. The maximum signal amplitude at all tempera-\ntures increaseswith increasing field. In addition, the neg-\nativeamplitude componentismostpronouncedat0.25T,\ndecreasing in amplitude at higher fields and vanishing en-\ntirely by 1 T. At this high field, we note that the photoin-\nduced magnetization dynamics look qualitatively similar\nto the photoinduced reflectivity signal ∆ R/Rshown in\nFig.2(a). In the ∆ R/Rdata, the measured signal is\nprimarily the result of Ti sublattice dynamics due to the\n1.9 eV intersite Ti-Ti excitation, and is dominated by Ti\nspin dynamics: the spin-lattice and spin relaxationterms\n(τs−landτslow). It follows that the MOKE signal mea-\nsured at 1 T is primarily a measure of Ti spin dynamics\ndue to its similarity with the ∆ R/Rsignal. Fits to the 1\nT MOKE data support this, yielding a component with\na timescale of 100 – 200 ps and a very similar tempera-\nture dependence when compared to τs−lextracted from\nthe ∆R/Rdata (see SM VI for details) [19]. As the field\nis lowered from 1 T, the magnetization dynamics must\nbe increasingly influenced by the Gd spins. The ferri-\nmagnetic nature of GTO, with two competing magnetic\nsublattices, is key to understanding the observed behav-\nior as we now discuss.6\n 50 \n 60 \n 80 \n100 \n150 \n250 10 \n 20 \n 25 \n 30 \n 35 \n 40 \nTime (ps)-2 -1 012 Residual (rad) 10 -5 Gd \nTi \n MZMZ > 0 MZ < 0 \nMZ > > 0 \nMMZ << 00\nMMZ >>>> 0\nT (K) (a)\n(b)\n0 10 20 30 40 50 H = 0.1 T\nH = 1 Tt < s-lt ≥ s-l \nt = 0\nHz\na-axis\nFIG. 5. (a) Schematic depiction of the spin dynamics. Pho-\ntoexcitation directly perturbs the Ti spins, which fluctuat e\nand decrease their projection along the applied field H (para l-\nlel toa-axis, +z) int < τs−lps, increasing the MOKE signal.\nAt longer times: if H and/or magnetic order is weak, induced\nspin fluctuations and the AFM exchange coupling lowers the\nprojection of Gd spins along z. Conversely, if H is larger\nthan the exchange field, at 1 T, Gd does not reorient and no\nnegative component of the signal is observed. (b) Coherent\nacoustic phonon response, isolated by subtracting the expo -\nnential fits from the time-resolved Kerr data (at 1 T applied\nfield). The dynamics appear similar to the ∆ R/Rresidual,\nimplying a common origin which we attribute to a coherent\nstrain wave launched by the pump pulse. The appearance\nof this signal in the Kerr response indicates coherent acous tic\nphonon manipulation of the magnetic order, presumably from\nexchange modulation.\nIV. DISCUSSION\nGTO is ferrimagnetic, the Ti and Gd sublattices cou-\npled via an AFM exchange interaction. Gd spins have a\nsignificantly larger magnetic moment than Ti, 7 µBvs1µBrespectively [15]. Below TC, at zero field, the two\nsublattices are aligned into fM domains such that there is\nnomacroscopicmoment. As the applied field Halongthe\na-axis is increased, spins are rotated to form long-range\ncollinearfMorder,withtheGdsublatticealignedparallel\nto H and Ti anti-parallel. In a field of only 0.1 T satu-\nration is approached, with spins slightly canted from the\na-axis/H and a net magnetization of M≈5µB. With in-\ncreasing field, canting and spin fluctuations are reduced,\nincreasing the net moment along H. As we approach 1 T,\nspin fluctuations are minimized and the magnetization\nbecomes saturated at M≈6.0µB[15]. The interac-\ntion of these two competing magnetic sublattices after\nphotoexcitation will depend on the temperature and ap-\nplied field and is illustrated in Fig. 5(a). The 1.9 eV\npump pulse directly excites the Ti sublattice, increas-\ning Ti spin fluctuations on timescales t < τs−l. This\ncauses partial reorientation and a decrease in the projec-\ntion of Ti spins along the Gd moment, corresponding to\na rapid increase of ∆ Mzand a rise in the MOKE signal\n(i.e. the Ti sublattice magnetization oriented along -zis\ndecreased, leading to an overall increase in the net mag-\nnetization in the + zdirection due to the ferrimagnetic\norder). Various pathways exist which may perturb the\nTi spins on such timescales, including spin-orbit coupling\n[50, 51] (orbital order is disrupted in t <8 ps, see SM III)\n[1, 11, 19, 40–45], and exchange modification, discussed\nbelow. Subsequently, energy is transferredto the Gd sub-\nlattice through spin-lattice thermalization on a timescale\nt≥τs−l. The spin-lattice coupling timescale of 100 –\n200 ps measured from fits to the ∆ R/Rand MOKE data\ncorresponds to the timescale on which the MOKE signal\nchanges sign, indicating the delayed contribution of Gd\nspins to the signal. Such ultrafast magnetic sublattice\ndynamics, albeit with different mechanisms, have been\ndiscussed in a variety of materials, including those with\nsimilar rare-earth/transition metal correlations [52, 53].\nThe behavior that follows is field-dependent. At low\nfields, at times on the order of τs−l, the additional heat\ntransfer and the strong AFM exchange coupling between\nGd spins and partially-reorientedTi spins causesa reduc-\ntion in the Gd moment along the field direction. This is\nseen as the negative component, decreasing the signal\non spin-lattice timescales until the net ∆ Mzis negative.\nAt higher field strengths, the applied field locks Gd mo-\nments in place parallel to the field direction, minimizing\nfluctuations. After photoexcitation, the net magnetiza-\ntion only increases as the anti-parallel Ti spins fluctuate\nand partially reorient. This is true also at low tempera-\ntureswherethemagneticorderismorefirmlyestablished,\nand explains why ∆ Mzgoes negative only in the weakly\nordered state near TC.\nFinally, we cannot discount the possibility of direct\nphoto-induced modification of the exchange interactions.\nWhile the simplest explanation of the MOKE signal in-\nvolves only heating and spin-lattice coupling, the over-\nall heating is small (no more than ∼4 K at the lowest\ntemperatures at the fluence used). It is therefore not un-7\n024\n0 50 100\nFrequency (GHz) 012FFT Amplitude (10 -7 )\n800 nm 10 \n 20 \n 25 \n 30 \n 35 \n 40 \n 50 \n 60 \n 80 \n100 \n295 T (K) \nFFT Amplitude (10 -6 )\n0 50 100012\nFrequency (GHz) 012\n40 ps 60 ps\n0 50 100\nFrequency (GHz) TC = 32 K\n0 100 200 300 \nTemperature (K) 01FFT (norm.) 4 GHz\n20 GHz\n50 GHz\n80 GHz(a)\n(b)(c)\n(d) (e)660 nm\nFIG. 6. FFT amplitude of the ∆ R/Rresidual, taken at pump/probe wavelength of 660 nm (a) and 80 0 nm (b). The red dashed\nlines indicate the approximate peak positions of the 660 nm F FT. Note the redshift to lower frequency at higher wavelengt h.\n(c) The integrated FFT amplitudes at various frequencies as a function of temperature, normalized. (d) The FFT limited t o\nthe first 40 ps and (e) 60 ps of the ∆ R/Rdata. The higher frequency mode emerges only after 40 ps.\nreasonable to consider more direct electronic changes to\nthe system. The exchange interaction in the titanates is\nhighly dependent on the Ti-O-Ti bond angle and degree\nof GdFeO 3distortion, as well as the orbital order and\noccupation [7, 11]. GTO in particular lies on the cusp of\nthe AFM-FM phase boundary, making it especially sus-\nceptible to changes in these parameters. Photoexcitation\ndirectly disrupts the orbital occupation, which could af-\nfect the octahedral distortion and thus the spin exchange\ninteraction. This in turn would provide the drive for re-\norientation of Ti spins and change in M zat timescales\nt < τs−l, and for subsequent perturbation of Gd spins\nthrough exchange coupling with Ti. Further calculations\nof the energy scales of the Ti-Gd exchange field and cor-\nresponding timescales are required to confirm this.\nTo compare the magnetic dynamics to the ∆ R/Rre-\nsponse, we fit the MOKE data to a series of exponentials\nsimilar to Eq. 1 and subtract the fits. Once again, a\nslow coherent oscillation is revealed, shown in Fig. 5(b)\nfor the data taken at 1 T. The similarity of the oscilla-\ntory Kerr signal to the oscillation in ∆ R/Ris striking –\nboth phonon modes have the same frequency, same time-\ndependent redshift, and same temperature dependence,\nwith the amplitude growing rapidly at TC. We rule out\nthe possibility of a magnon – at lower fields there is no\nchange in the frequency of oscillation as we would ex-\npect from coherent spin precession (SM VII) [19]. The\namplitude is highly field-dependent however, becoming\nmuch smaller at lower fields. These observations suggestthat the oscillatory mode in the MOKE signal, neces-\nsarily a magnetic phenomenon due to the nature of the\nmeasurement technique, has the same origin as the os-\ncillatory mode in the ∆ R/Rsignal. This is consistent\nwith our interpretation of an acoustic strain wave with\nstrong magneto-elastic coupling. This mechanism has\nbeen studied in a variety of ferromagnetic systems, and\ninvolves elastic stress modifying the magnetic anisotropy,\nwhich exerts a torque on the spins and alters the net\nmagnetization [54–56].\nTo quantify the acoustic phonon response, we show in\nFig.6(a) the FFT of the full ∆ R/Rresidual, taken from\nFig.3(b) (inset). The oscillatory mode with ∼20 ps pe-\nriod featured in Fig. 3(b) appears as a strong peak at\n∼50 GHz. In this region of interest, it is apparent that\nthere are additional higher frequency modes in addition\nto the 50 GHz mode. The temperature dependence is\nalso clear; while the FFT amplitude is nearly constant\nat high temperatures, it grows rapidly upon approaching\nTC= 32 K and a higher frequency peak at ∼80 GHz\nemerges. This again suggests coupling to the magnetic\norder. Fig. 6(b) applies the same FFT analysis to data\ntaken at an increased pump/probe wavelengthof 800 nm.\nThe features are similar, but exhibit a clear redshift as\nindicated by the red dashed lines. This behavior is con-\nsistent with an acoustic strain wave since it arises (for\n∆R/R) from interference of the probe with itself. The\nphonon frequency is wavelength dependent, its form is8\ngiven by:\nf= 2nv/λ, (2)\nwherenis the index of refraction, vis the sound veloc-\nity, and λis the probe wavelength [49]. As we observe,\na higher probe wavelength results in a lower frequency\nacoustic phonon. Using the measured index of refraction\nnin Fig.1(b) wecanalsoestimatethe sound velocity. At\nthe lower frequency peak near 50 GHz we obtain a sound\nvelocity of 7 .2×103m/sand 7.7×103m/sfor a 660 and\n800 nm probe, respectively. This is a very reasonable\nrange for acoustic propagation in solid materials. These\nresults, and the fact that the oscillation frequency does\nnot depend on magnetic field, confirms our classification\nof the phonon mode as an acoustic strain wave.\nTomorecloselyexaminethelinktomagnetism,weplot\nthe integrated FFT amplitudes for all frequency peaks in\nFig.6(c). The normalized curves show a striking trend;\nthe amplitude is nearly constant at high temperatures,\nbut sharply increases at or very near to the magnetic or-\ndering transition. The temperature dependence of the\nFFT amplitudes follows the magnetic order parameter\nandisremarkablysimilartothedivergenceoneexpectsat\na second-order magnetic phase transition. This indicates\nthe presence of magneto-elastic coupling. The acoustic\nattenuation of sound waves near magnetic phase tran-\nsitions is well studied, and literature suggests that the\nattenuation follows power law behavior, similar to our re-\nsult [54]. In the vicinity of TC, energy density and spin\nfluctuations play the primary role in attenuation. This\nbehavior has been studied in a wide range of magneto-\nelasticallycoupledmaterials, including Ni [54], CoF 2[57],\nand MnF 2[58].\nA final interesting feature to note is shown in Fig. 6(d-\ne), comparing an FFT of the ∆R/Rdata limited to the\nfirst 40 ps (d) and to the first 60 ps (e) of the scan. This\nanalysis reveals that the high frequency component at\n80 GHz begins to emerge only after 40 ps, which is also\nvisible in the time-domain data (Fig. 3(b) inset). This\ntimescale is similar to the spin-lattice coupling timescale\nmeasured in both ∆ R/Rand MOKE, which ranges from\n∼50 – 150 ps. We have also discussed the spin dynam-\nics following photoexcitation, where Ti spins are imme-\ndiately perturbed and Gd follows after exchange path-\nway alterations and spin-lattice thermalization. Given\nthe similar timescales, we suggest that the emergence of\nthe 80 GHz mode indicates the onset of Gd spin dynam-\nics. Roughly 50 ps after photoexcitation the Gd spin sub-\nsystem begins thermalizing and fluctuating. This damps\nthe acoustic oscillation and changes the magnetic back-\nground. The nowhigherenergyofthe Gd spins altersthe\nspin-phonon and magnetostrictive interaction strengths,\nresulting in a change to the magnetically-coupled elas-\ntic parameters of the lattice and a subsequent shift in\nphonon frequency.\nA microscopic description of magneto-elastic coupling\ninvolves a transient modification of the exchange interac-\ntion. As the acoustic wave propagates it modulates thedistance between lattice sites and spins. This in turn pro-\nduces a periodic modification of the exchange interaction\nbetween neighboring spins, coupling the acoustic wave to\nthe magnetic order parameters. The result is an attenua-\ntion of the acoustic wave in the high-temperature phase\nwhere spin fluctuations are large, lessening as spin cor-\nrelations increase in the low temperature ordered phase.\nThe same mechanism decreases acoustic attenuation, in-\ncreasing the phonon amplitude, in an applied magnetic\nfield as observed in our MOKE signal. This has been\ndescribed by an approximate analytical theory [58–60],\nwhich generally predicts maximal acoustic damping at\nthe critical point and a MHz frequency shift in the or-\ndered phase. We observe that the damping is consis-\ntently large throughout the high temperature paramag-\nnetic phase, and we do not observe such a frequency shift\nwith temperature. In our experiment, however, a MHz\nfrequency shift is too small to be observed, and the dy-\nnamics at picosecond timescales are strongly coupled to\nout-of-equilibrium degrees of freedom that will affect the\nacousticwavepropagationandattenuationinotherunan-\nticipated ways.\nThephonon behaviorweobserveundoubtedly suggests\na strong coupling of the lattice to the magnetic order\nin GdTiO 3. Furthermore, the mechanism implies tran-\nsient exchange modification on an ultrafast timescale.\nThese conclusions are not without precedence. Ultra-\nfast magneto-elastic coupling has been demonstrated by\nBigotet al., for example, in Ni thin films [61], with exper-\niments going so far as to control the magnetic precession\nthrough acoustic pulses [62]. Kimel et al.have shown\noptical quenching of magnetic order through phonon-\nmagnon coupling in FeBO 3[48] and Nova et al.have\nshownthatMid-IRandTHzexcitationresonantwithspe-\ncific lattice modes is able to drive collective spin preces-\nsion [63]. Our work represents another potential method\nof using light to indirectly alter the magnetic degrees of\nfreedom on ultrafast timescales, through coupling to an\nacoustic phonon mode.\nV. CONCLUSION\nWe have used a multi-modal approach, consisting\nof time-resolved photoinduced reflectivity and magneto-\noptical Kerr (MOKE) spectroscopy, to study magneto-\nelastic coupling in the ferrimagnetic insulator GdTiO 3.\nWe observe multiple, clear signatures of the ferrimag-\nnetically ordered phase at TC= 32 K in both signals,\nand measure spin-lattice thermalization timescales τs−l\non the order of 100 picoseconds, as might be expected in\na magnetic insulator.\nFrom the MOKE signal we observe long-lived spin dy-\nnamics and optical perturbation of the ferrimagnetic or-\nder. This includes a change in sign of the photoinduced\nmagnetization on the same timescale as spin-lattice cou-\npling. The ferrimagnetic nature of GTO, with two mag-9\nnetic sublattices coupled antiferromagnetically, is respon-\nsible. Photoexcitation at 660 nm directly perturbs the\nTi moments, increasing fluctuations and causing a par-\ntial reorientation and decrease in the projection of Ti\nspins along the Gd moment. This is measured as an\nincrease in the MOKE signal. Heat is then transferred\ntothe Gdsubsystem throughspin-latticecoupling, which\nwhen combinedwith the AFM exchangeinteractionleads\nto a reduction of the Gd magnetic moment along the z-\ndirection, lowering the net magnetization. Modified ex-\nchange pathways likely also play a role in the delayed\nreorientation of Gd spins on these timescales. The data\nshows that (a) there is a delayed response of the Gd ions\ntotheopticalexcitationand(b)thatspin-latticecoupling\nand the AFM exchange interaction facilitates this.\nIn both the reflectivity and MOKE signals, a clear\ncoherent acoustic phonon is present. This strain wave\nlaunched by pump is intimately tied to the sample mag-\nnetism, with an amplitude that grows sharply at TCand\nclosely follows the magnetic order parameter. As the\nacoustic wave propagates it periodically alters the dis-\ntance between local spins, modifying the exchange inter-\naction. In this way, the lattice parameters are coupled to\nthe magnetic order, which causes an attenuation of the\nacoustic mode near and above TC, where spin fluctua-\ntions are large. This represents a laser-induced modifica-\ntion of the exchange interaction on ultrafast timescales\nthroughcouplingtoanacousticphononmode. Whilethe-\nory exists to describe magneto-elastic coupling, it is not\nparticularly well-suited to the experiment and timescalesmeasured here. A deeper theoretical understanding of\nthe mechanisms at work would be instrumental in quan-\ntifying our results and motivating further studies. This\nwork also suggests that more controlled excitation may\nbe of interest in transiently controlling the properties of\nmaterials. An experiment ofthis naturehas alreadybeen\nproposed to modify the exchange interaction in GTO, us-\ning a resonant mid-IR pulse to directly excite specific\nphonon modes [17]. The work performed here indicates\nthe potential for GTO, and likely other titanates, as tun-\nable magnetic materials, and highlights the need for fur-\nther investigations of this nature on the road to coherent\ncontrol of materials on ultrafast timescales.\nVI. ACKNOWLEDGEMENTS\nWe thank Leon Balents for helpful discussions and\nassistance with interpretation of the data. This work\nwas supported primarily by ARO Award W911NF-16-\n1-0361 and additional support was provided by the W\nM Keck Foundation (SDW). The MRL Shared Exper-\nimental Facilities used for sample characterization are\nsupported by the MRSEC Program of the NSF under\nAward No. DMR 1720256; a member of the NSF-funded\nMaterials Research Facilities Network. The work at\nHYU was supported by the Basic Science Research Pro-\ngram through the National Research Foundation of Ko-\nrea (NRF), funded by the Ministry of Science, ICT and\nFuture Planning (2019R1A2C1084237).\n[1]M. Mochizuki and M. Imada, Orbital physics in the per-\novskite Ti oxides, New J. Phys. 6, 154 (2004) .\n[2]Y. Tokura, Fillingness dependence of electronic struc-\ntures in strongly correlated electron systems: Titanates\nand vanadates, J. Phys. Chem. Solids 53, 1619 (1992) .\n[3]C. W. Turnerand J. Greedan, Ferrimagnetism in the rare\nearth titanium (III) oxides, RTiO 3; R = Gd, Tb, Dy, Ho,\nEr, Tm, J. Solid State Chem. 34, 207 (1980) .\n[4]Y. Okimoto, T. Katsufuji, Y. Okada, T. Arima, and\nY. Tokura, Optical spectra in (La,Y)TiO 3: Variation of\nMott-Hubbard gap features with change of electron cor-\nrelation and band filling, Phys. Rev. B 51, 9581 (1995) .\n[5]M. Itoh, M. Tsuchiya, H. Tanaka, and K. Motoya, Or-\nbital Ordering and Local Magnetic Properties of Mott-\nHubbard Insulators YTiO 3and LaTiO 3: NMR Study, J.\nPhys. Soc. Japan 68, 2783 (1999).\n[6]M. Mochizuki and M. Imada, Origin of G-type anti-\nferromagnetism and orbital-spin structures in LaTiO 3,\nJ. Phys. Soc. Japan 70, 2872 (2001) .\n[7]M. Mochizuki and M. Imada, Magnetic Phase\nTransition of the Perovskite-Type Ti Oxides,\nJ. Phys. Soc. Japan 69, 1982 (2000) .\n[8]H. D. Zhou and J. B. Goodenough, Localized\nor itinerant TiO 3electrons in RTiO 3perovskites,\nJ. Phys. Condens. Matter 17, 7395 (2005) .[9]K. Takubo, M. Shimuta, J. E. Kim, K. Kato, M. Takata,\nand T. Katsufuji, Crossover behavior of the crystal struc-\nture and the relation to magnetism in perovskite RTiO 3,\nPhys. Rev. B 82, 020401(R) (2010) .\n[10]E. Pavarini, S. Biermann, A. Poteryaev, A. I. Lichten-\nstein, A. Georges, and O. K. Andersen, Mott Transition\nand Suppression of Orbital Fluctuations in Orthorhom-\nbic 3d1Perovskites, Phys. Rev. Lett. 92, 176403 (2004) .\n[11]M. Mochizuki and M. Imada, Magnetic and Or-\nbital States and Their Phase Transition of the\nPerovskite-Type Ti Oxides: Strong Coupling Approach,\nJ. Phys. Soc. Japan 70, 1777 (2001) .\n[12]A. C. Komarek, H. Roth, M. Cwik, W. D. Stein, J. Baier,\nM. Kriener, F. Bour´ ee, T. Lorenz, and M. Braden, Mag-\nnetoelastic coupling in RTiO 3(R=La,Nd,Sm,Gd,Y) in-\nvestigated with diffraction techniquesand thermal expan-\nsion measurements, Phys. Rev. B 75, 224402 (2007) .\n[13]J. Y. Zhang, C. A. Jackson, S. Raghavan, J. Hwang,\nand S. Stemmer, Magnetism and local struc-\nture in low-dimensional Mott insulating GdTiO 3,\nPhys. Rev. B 88, 121104(R) (2013) .\n[14]J. Varignon, M. N. Grisolia, D. Preziosi, P. Ghosez, and\nM. Bibes, Origin of the orbital and spin ordering in rare-\nearth titanates, Phys. Rev. B 96, 235106 (2017) .\n[15]G. Amow, J. S. Zhou, and J. B. Goodenough, Pe-\nculiar magnetism of the Sm (1−x)GdxTiO3system,10\nJ. Solid State Chem. 154, 619 (2000) .\n[16]J. Liu, K. Hejazi, and L. Balents, Flo-\nquet Engineering of Multiorbital Mott Insula-\ntors: Applications to Orthorhombic Titanates,\nPhys. Rev. Lett. 121, 107201 (2018) .\n[17]G. Khalsa and N. A. Benedek, Ultrafast optically in-\nduced ferromagnetic/anti-ferromagnetic phase transitio n\nin GdTiO 3from first principles, npj Quantum Mater. 3,\n10.1038/s41535-018-0086-3 (2018).\n[18]M. Gu and J. M. Rondinelli, Nonlinear phononic control\nand emergent magnetism in Mott insulating titanates,\nPhys. Rev. B 98, 024102 (2018) .\n[19]See Supplemental Material at [URL will be inserted by\npublisher] for additional measurements and supporting\nanalysis.\n[20]P. Moetakef, J. Y. Zhang, S. Raghavan, A. P. Ka-\njdos, and S. Stemmer, Growth window and effect\nof substrate symmetry in hybrid molecular beam\nepitaxy of a Mott insulating rare earth titanate,\nJ. Vac. Sci. Technol. A Vacuum, Surfaces, Film. 31, 041503 (2013) .\n[21]J. L. Schmehr, M. Aling, E. Zoghlin, and S. D.\nWilson, High-pressure laser floating zone furnace,\nRev. Sci. Instrum. 90, 043906 (2019) .\n[22]A. B. Kuzmenko, Kramers-Kronig con-\nstrained variational analysis of optical spectra,\nRev. Sci. Instrum. 76, 083108 (2005) .\n[23]R.C. Jones, ANewCalculus for theTreatmentofOptical\nSystems I. Description and Discussion of the Calculus,\nJ. Opt. Soc. Am. 31, 488 (1941) .\n[24]M. Veis, Optical interactions in thin films of selected mag-\nnetic oxides , Doctoral thesis, Charles University, Prague\n(2009).\n[25]D. S. Kliger, J. W. Lewis, and C. E. Randall, Intro-\nduction to the Jones Calculus, Mueller Calculus, and\nPoincare Sphere,in Polariz. Light Opt. Spectrosc. ,edited\nby D. S. Kliger, J. W. Lewis, and C. E. Randall (Aca-\ndemic Press, Boston, 1990) pp. 59–101.\n[26]M. Nyvlt, Optical interactions in ultrathin magnetic film\nstructures , Thesis, Charles University, Prague (1996).\n[27]J. Wang, Time-Resolved Magneto-Optical Spectroscopy,\ninOpt. Tech. Solid-State Mater. Charact. , edited by R. P.\nPrasankumar and A. J. Taylor (CRC Press, Boca Raton,\nFL, 2012).\n[28]D. Crandles, T. Timusk, J. Garrett, and J. Greedan,\nThe midinfrared absorption in RTiO 3perovskites (R\n= La, Ce, Pr, Nd, Sm, Gd): The Hubbard gap?,\nPhys. C Supercond. 201, 407 (1992) .\n[29]L. Bjaalie, A. Verma, B. Himmetoglu, A. Jan-\notti, S. Raghavan, V. Protasenko, E. H. Steenber-\ngen, D. Jena, S. Stemmer, and C. G. Van de Walle,\nDetermination of the Mott-Hubbard gap in GdTiO 3,\nPhys. Rev. B 92, 085111 (2015) .\n[30]P. C. Hohenberg and B. I. Halperin, Theory of dynamic\ncritical phenomena, Rev. Mod. Phys. 49, 435 (1977) .\n[31]F. Wang, N. Hatano, and M. Suzuki, Study\non dynamical critical exponents of the Ising\nmodel using the damage spreading method,\nJ. Phys. A. Math. Gen. 28, 4543 (1995) .\n[32]A. Pelissetto and E. Vicari, Critical phe-\nnomena and renormalization-group theory,\nPhys. Rep. 368, 549 (2002) .\n[33]H. E. Stanley, Introduction to phase transitions and crit-\nical phenomena (Oxford University Press, 1987).[34]M. Campostrini, M. Hasenbusch, A. Pelissetto, P. Rossi,\nand E. Vicari, Critical exponents and equation of state\nof the three-dimensional Heisenberg universality class,\nPhys. Rev. B 65, 144520 (2002) .\n[35]J. P. Hinton, S. Patankar, E. Thewalt, A. Ruiz,\nG. Lopez, N. Breznay, A. Vishwanath, J. Analytis,\nJ. Orenstein, J. D. Koralek, and I. Kimchi, Photoex-\ncitedstates oftheharmonic honeycombiridate γ-Li2IrO3,\nPhys. Rev. B 92, 115154 (2015) .\n[36]P. Moetakef, D. G. Ouellette, J. Y. Zhang, T. A. Cain,\nS. J. Allen, and S. Stemmer, Growth and properties of\nGdTiO 3films prepared by hybrid molecular beam epi-\ntaxy,J. Cryst. Growth 355, 166 (2012) .\n[37]J. Qi, L. Yan, H. D. Zhou, J. X. Zhu, S. A. Trug-\nman, A. J. Taylor, Q. X. Jia, and R. P. Prasanku-\nmar, Coexistence of coupled magnetic phases in epitaxial\nTbMnO 3films revealed by ultrafast optical spectroscopy,\nAppl. Phys. Lett. 101, 122904 (2012) .\n[38]S. Wall, D. Prabhakaran, A. T. Boothroyd, and A. Caval-\nleri, Ultrafast Coupling between Light, Coherent Lattice\nVibrations, and the Magnetic Structure of Semicovalent\nLaMnO 3,Phys. Rev. Lett. 103, 097402 (2009) .\n[39]K. Miyasaka, M. Nakamura, Y. Ogimoto,\nH. Tamaru, and K. Miyano, Ultrafast photoin-\nduced magnetic moment in a charge-orbital-\nordered antiferromagnetic Nd 0.5Sr0.5MnO3thin film,\nPhys. Rev. B 74, 012401 (2006) .\n[40]Y. Furukawa, I. Okamura, K. Kumagai, Y. Taguchi, and\nY.Tokura, NMRStudyonElectronic andMagnetic State\nin RTiO 3(R=La, Y), J. Low Temp. Phys. 105, 413\n(1996).\n[41]Y. Furukawa, I. Okamura, K. Kumagai, Y. Taguchi,\nand Y. Tokura, NMR study of ferromag-\nnetic YTiO 3and antiferromagnetic LaTiO 3,\nPhys. B Condens. Matter 237-238 , 39 (1997) .\n[42]M. Itoh and M. Tsuchiya, Orbital or-\ndering in YTiO 3observed by NMR,\nJ. Magn. Magn. Mater. 226-230 , 874 (2001) .\n[43]S. Tomimoto, S. Miyasaka, T. Ogasawara, H. Okamoto,\nand Y. Tokura, Ultrafast photoinduced melting of orbital\norder in LaVO 3,Phys. Rev. B 68, 035106 (2003) .\n[44]A. G¨ ossling, R. Schmitz, H. Roth, M. W. Haverkort,\nT. Lorenz, J. A. Mydosh, E. M¨ uller-Hartmann,\nand M. Gr¨ uninger, Mott-Hubbard exciton in\nthe optical conductivity of YTiO 3and SmTiO 3,\nPhys. Rev. B 78, 075122 (2008) .\n[45]F. Novelli, D. Fausti, J. Reul, F. Cilento, P. H. M.\nvan Loosdrecht, A. A. Nugroho, T. T. M. Palstra,\nM. Gr¨ uninger, and F. Parmigiani, Ultrafast optical spec-\ntroscopy of the lowest energy excitations in the Mott in-\nsulator compound YVO 3: Evidence for Hubbard-typeex-\ncitons,Phys. Rev. B 86, 165135 (2012) .\n[46]J. Wang, C. Sun, J. Kono, A. Oiwa, H. Munekata,\nL. Cywi´ nski, and L. J. Sham, Ultrafast Quenching of Fer-\nromagnetism in InMnAs Induced by Intense Laser Irra-\ndiation, Phys. Rev. Lett. 95, 167401 (2005) .\n[47]A. Vaterlaus, T. Beutler, and F. Meier, Spin-lattice\nrelaxation time of ferromagnetic gadolinium deter-\nmined with time-resolved spin-polarized photoemission,\nPhys. Rev. Lett. 67, 3314 (1991) .\n[48]A. V. Kimel, R. V. Pisarev, J. Hohlfeld, and T. Rasing,\nUltrafast Quenching of the Antiferromagnetic Order in\nFeBO 3: Direct Optical Probing of the Phonon-Magnon\nCoupling, Phys. Rev. Lett. 89, 287401 (2002) .11\n[49]C. Thomsen, H. T. Grahn, H. J. Maris, and J. Tauc, Sur-\nface generation and detection of phonons by picosecond\nlight pulses, Phys. Rev. B 34, 4129 (1986) .\n[50]E. Beaurepaire, M. Maret, V. Halt´ e, J. C. Merle,\nA. Daunois, and J. Y. Bigot, Spin dynamics in CoPt 3\nalloy films: A magnetic phase transition in the femtosec-\nond time scale, Phys. Rev. B 58, 12134 (1998) .\n[51]T. Ogasawara, K. Ohgushi, Y. Tomioka, K. S. Taka-\nhashi, H. Okamoto, M. Kawasaki, and Y. Tokura,\nGeneral features of photoinduced spin dynam-\nics in ferromagnetic and ferrimagnetic compounds,\nPhys. Rev. Lett. 94, 087202 (2005) .\n[52]I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius,\nH. A. D¨ urr, T. A. Ostler, J. Barker, R. F. L. Evans, R. W.\nChantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk, T. Ras-\ning, and A. V. Kimel, Transient ferromagnetic-like state\nmediating ultrafast reversal of antiferromagnetically co u-\npled spins, Nature472, 205 (2011) .\n[53]Z. Chen, S. Li, S. Zhou, and T. Lai, Ultra-\nfast dynamics of 4f electron spins in TbFeCo film\ndriven by inter-atomic 3d5d4f exchange coupling,\nNew J. Phys. 21, 123007 (2019) .\n[54]M. Weiler, L. Dreher, C. Heeg, H. Huebl, R. Gross,\nM. S. Brandt, and S. T. B. Goennenwein, Elastically\nDriven Ferromagnetic Resonance in Nickel Thin Films,\nPhys. Rev. Lett. 106, 117601 (2011) .\n[55]E. Rossi, O. G. Heinonen, and A. H. MacDonald, Dynam-\nics of magnetization coupled to a thermal bath of elasticmodes,Phys. Rev. B 72, 174412 (2005) .\n[56]S. Streib, H. Keshtgar, and G. E. W. Bauer, Damp-\ning of Magnetization Dynamics by Phonon Pumping,\nPhys. Rev. Lett. 121, 027202 (2018) .\n[57]R. I. Thomson, T. Chatterji, and M. A. Carpen-\nter, CoF 2: A model system for magnetoelastic cou-\npling and elastic softening mechanisms associated\nwith paramagnetic antiferromagnetic phase transitions,\nJ. Phys. Condens. Matter 26, 146001 (2014) .\n[58]J. R. Neighbors and R. W. Moss, Ultrasonic at-\ntenuation near the magnetic critical point of MnF 2,\nPhys. Rev. 173, 542 (1968) .\n[59]H. S. Bennett and E. Pytte, Ultrasonic Attenuation in\nthe Heisenberg Paramagnet, Phys. Rev. 155, 553 (1967).\n[60]S. K. Ghatak, Acoustic Attenuation and Frequency Shift\nin Ferromagnetic Insulators at Low Temperature, Phys.\nRev. B5, 3702 (1972).\n[61]J. W. Kim, M. Vomir, and J. Y. Bigot, Ul-\ntrafast Magnetoacoustics in Nickel Films,\nPhys. Rev. Lett. 109, 166601 (2012) .\n[62]J. W. Kim, M. Vomir, and J. Y. Bigot, Controlling the\nspins angular momentum in ferromagnets with sequences\nof picosecond acoustic pulses, Sci. Rep. 5, 1 (2015) .\n[63]T. F. Nova, A. Cartella, A. Cantaluppi, M. F¨ orst,\nD. Bossini, R. V. Mikhaylovskiy, A. V. Kimel, R. Merlin,\nand A. Cavalleri, An effective magnetic field from opti-\ncally driven phonons, Nat. Phys. 13, 132 (2017) ." }, { "title": "2401.05130v3.Laser_induced_ultrafast_Gd_4f_spin_dynamics_at_the_surface_of_amorphous_CoxGd100_x_ferrimagnetic_alloys.pdf", "content": "1 \n Laser induced ultrafast Gd 4f spin dynamics at the surface of amorphous Co xGd 100-x \nferrimagnetic alloys. \n \nM. Pacé 1, D. Gupta 1, T. Ferté 1, M. Riepp 1, G. Malinowski 2, M. Hehn 2, F. Pressacco 3, M. \nSilly 3, F. Sirotti 4, C. Boeglin 1 and N. Bergeard 1 \n1 Université de Strasbourg, CNRS, Institut de Physiqu e et Chimie des Matériaux de Strasbourg, \nUMR 7504, F-67000 Strasbourg, France. \n2 Institut Jean Lamour, CNRS - Université de Lorraine , 54011 Nancy, France. \n3 Synchrotron SOLEIL, L'Orme des Merisiers, Saint-Aub in, 91192 Gif-sur-Yvette, France \n4 Physique de la Matière Condensée, Ecole Polytechni que, CNRS, 91128 Palaiseau, France \n \nAbstract : \nWe have investigated the laser induced ultrafast dy namics of Gd 4f spins at the surface \nof Co xGd 100-x alloys by means of surface-sensitive and time-reso lved dichroic resonant Auger \nspectroscopy. We have observed that the laser induc ed quenching of Gd 4f magnetic order at \nthe surface of the Co xGd 100-x alloys occur on a much longer time scale than that previously \nreported in “bulk sensitive” time-resolved experime nts. In parallel, we have characterized the \nstatic structural and magnetic properties at the su rface and in the bulk of these alloys by \ncombining Physical Property Measurement System (PPM S) magnetometry with X-ray \nMagnetic Circular Dichroism in absorption spectrosc opy (XMCD) and X-Ray Photoelectron \nspectroscopy (XPS). The PPMS and XMCD measurements give information regarding the \ncomposition in the bulk of the alloys. The XPS meas urements show non-homogeneous \ncomposition at the surface of the alloys with a str ongly increased Gd content within the first \nlayers compared to the nominal bulk values. Such la rger Gd concentration results in a reduced \nindirect Gd 4f spin-lattice coupling. It explains t he “slower” Gd 4f demagnetization we have \nobserved in our surface-sensitive and time-resolved measurements compared to that previously \nreported by “bulk-sensitive” measurements. \nIntroduction : 2 \n The discovery of deterministic helicity-independent all-optical spin switching (HI-\nAOS) [OST12] has driven intensive investigations on the laser induced ultrafast spin dynamics \nin ferrimagnetic RE-TM alloys [RAD11, LOP13, GRA13, BER14, RAD15, HIG16, HEN19]. \nThe element- and time-resolved experiments have sho wn that the RE 4f and TM 3d spin \ndynamics occur on distinct time-scales, which is co nsidered as the key ingredient for HI-AOS \n[IIH18]. They have also shown that the characterist ic times associated with these ultrafast \ndynamical processes depend on the temperature and o n the alloy composition [LOP13, FER17, \nCHE19, REN20, FER21, FER23]. Recently, Hennecke et al have also reported on transient \nmagnetization in-depth gradients that appear in GdF e alloys excited by femtosecond laser \npulses [HEN22]. These seminal works call for furthe r systematic experimental and theoretical \ninvestigations to reveal the correlations between t he static magnetic properties of these \nferrimagnetic alloys and their unusual laser induce d ultrafast spin dynamics [ATX14]. Such \nknowledge is of prime importance to identify the mo st appropriate materials for technological \napplications. In parallel, previous experimental wo rks have revealed that amorphous RE-TM \nferrimagnetic alloys show lateral [GRA13] and in de pth non-homogeneous composition \n[HEB16, HAL18, INY23] as well as RE segregation at interfaces [SHE81, BER17]. On the \nbasis of these previous observations, one can legit imately wonder whether the laser induced RE \n4f spin dynamics at the surface of the alloy is aff ected by the composition discrepancies in \nrespect with that in the bulk [FER23]. Disparate sp in dynamics at the surface and in the bulk \nwould be detrimental for technological applications , such as data storage devices that require \nthin magnetic layers [SEN21]. \n In this work, we have investigated the laser induc ed ultrafast dynamics of Gd 4f spins \nat the surface of Co xGd 100-x alloys for bulk compositions x = 80 and x = 65 by means of surface-\nsensitive time-resolved dichroic resonant Auger spe ctroscopy (TR-DRAS) [BEA13, SIL17]. \nWe have observed that the laser induced quenching o f Gd 4f magnetic order at the surface of \nthe Co xGd 100-x alloys occur on a much longer time scale than that previously reported in bulk \nsensitive time-resolved experiments [LOP13, RAD15, FER23]. In parallel, we have \ncharacterized the structural and magnetic static pr operties at the surface and in the bulk of these \nalloys by combining Physical Property Measurement S ystem (PPMS) magnetometry with X-\nray Magnetic Circular Dichroism in absorption spect roscopy (XMCD) [NAK99] and X-Ray \nPhotoelectron spectroscopy (XPS). The PPMS measurem ents confirm the nominal bulk \ncomposition of the alloys. The XPS measurements evi dence the non-homogeneous composition \nat the surface of the alloys with a strongly increa sed Gd content within the first layers [SHE81, 3 \n BER17, BAL16, HAS18, INY23]. As a consequence, we i nfer that the larger Gd concentration \nin these top layers compared to that of the nominal bulk composition allows to explain the \n“slower” Gd 4f demagnetization we have observed in our surface-sensitive and time-resolved \nmeasurements compared to that previously obtained w ith “bulk-sensitive” techniques [FER23]. \nExperimental : (methods, techniques, and materials studied) \nThe Co xGd 100-x(20) alloy layers were grown by DC co-sputtering on \n[Ta(5)/Cu(20)/Ta(5)] x5 multilayers deposited on Si substrates (units in n m). The layers were \ncapped with an Al(5) layer to prevent degradation o f the alloy in the ambient atmosphere \n[BER17]. This Al(5) layer is partly oxidized during exposure to air, but its thickness ensures \nthat a metallic Al layer remains in contact with th e Co xGd 100-x(20) alloys. Therefore, the actual \nstructure of the samples is Al 2O3(~3)/Al(~2)/Co xGd 100-x(20)/ [Ta(5)/Cu(20)/Ta(5)] 5/Si. It is \nworth mentioning that the samples were deposited on a 1” Si wafer which was cleaved after \ndeposition to ensure identical samples for the vari ous experimental techniques. In this study, \nwe have investigated the static and dynamics magnet ic properties of alloys with the nominal \ncomposition Co 80 Gd 20 and Co 65 Gd 35 respectively. The bulk compositions of the alloys w ere \ncontrolled by tuning the deposition rates of the cr ucibles. The actual alloy compositions were \nverified by recording the temperature dependence of magnetization by mean of Physical \nProperty Measurement System (PPMS) cryostat with Vi brating Sample Magnetometer (VSM) \noption head (figure 1). Indeed, we have compared th e temperature of magnetic compensation \nor the Curie temperature extracted from these measu rements to the tabulated values in literature \n[TAO74, HAN89] to estimate the composition. \nThe CoGd magnetic properties were also characterize d by mean of X-ray Magnetic \nCircular Dichroism in absorption spectroscopy (XMCD ) [NAK99]. We have recorded the X-\nray Absorption Spectra (XAS) at the Gd M 4,5 edges in the total electron yield acquisition mode \n(figure 3). These measurements were performed in th e main UHV chamber on the TEMPO \nbeamline at synchrotron SOLEIL by using circularly polarized soft X-ray [POL10]. The X-ray \nbeam impinges the sample at an angle of αX = 44° in respect with the samples normal (figure \n2) since the Co xGd 100-x alloys display in-plane magnetic anisotropy with a preferential axis \n[TAY76, BER17]. The data acquisitions were performe d in the remnant magnetic state after \nsaturation along the in-plane preferential axis by a magnetic pulse of 300 ms duration and 200 \nOe maximum amplitude. The XMCD spectra were obtaine d by recording the XAS spectra with \nopposite directions of the external magnetic field (figure 3a and 3c) [BER17]. The coercive \nfield of the Co 65 Gd 35 alloy was below 200 Oe for temperatures (T) ranging between 80 K to 4 \n 300 K, while the coercive field of the Co 80 Gd 20 alloy was below 200 Oe for T > 200K. Since \nwe have restricted ourselves to qualitative analysi s, we haven’t corrected the displayed spectra \nfor the saturation effects [NAK99]. Prior to XMCD m easurements, the alloys were sputtered in \nthe preparation chamber of the TEMPO beamline to pa rtly remove the Al capping layer to \nincrease the signal. We have kept a metallic Al lay er to prevent the oxidization during the \nmeasurements [BER17]. \nAmorphous RE-TM ferrimagnetic alloys are known to d isplay in-depth compositional \ngradients [HEB16, HAL18, INY23] or RE segregation a t the surface [SHE81, BER17]. The \ncomposition at the surface of the alloys was thus i nvestigated by mean of X-Ray Photoelectron \nSpectroscopy (XPS). The measurements were performed in the main chamber of the TEMPO \nbeamline by using the SES2002 photoelectron analyze r. Prior to XPS measurements, the alloys \nwere sputtered in-situ with Ar+ ions to completely remove the Al 2O3(~3)/Al(~2) capping layer. \nHowever, we have regularly monitored the thickness of the remaining capping layer as a \nfunction of sputtering time in order to minimize th e etching of the alloys. We have set the \nphoton energy to 700 eV and we have recorded the XP S spectra of the Al 2p, Gd 4f, Gd 4d as \nwell as the Co 3d and Co 3p core-levels (figure 4a and 4b). The sample’s normal was aligned \nwith the entrance of the photoelectron analyzer whi le the X-ray beam impinged the samples at \nan angle of αX = 44° in respect with the sample’s normal (figure 2). We have repeated these \nmeasurements at photon energies of 400 eV and 1000 eV. This protocol allows varying the \nsurface sensitivity of core-level XPS since the ele ctron inelastic mean free path depends on the \nkinetic energy of photoelectrons and thus on the ph oton energy (table 1) [JAB11]. As a \nconsequence, the larger photon energies allow probi ng deeper into the layer. In figure 5, we \nshow the Gd 4f and Co 3p core-levels as a function of the photon energy for the Co 80 Gd 20 (a) \nand Co 65 Gd 35 (b) alloys, respectively. A Shirley-like background was subtracted from the \nexperimental XPS core-level peaks, while the peak’s area was normalized by the \nphotoionization cross-section of the materials [YEH 85]. We have then normalized the spectra \nwith the height of the Gd 4f peak for direct compar ison. In order to quantify the Gd excess at \nthe surface of the alloy in respect with the nomina l composition, we define R as the ratio \nbetween the Gd 4f and Co 3p core-level peak’s area. The dependence of R on the photon energy \nis depicted in figure 6a. We have developed an elem entary model to provide a qualitative \ndescription of the alloy profiles [PAC23]. We have considered that the XPS signal of the specie \nx coming from the i th layer /g1827/g3051/g3036 is given by /g1857/g1876/g1868 /g4672 /uni2211 /g4672/g2879/g2869\n/g3030/g3042/g3046 /g4666/g3087/g4667./g4666/g3030/g3004/g3042 /g4666/g3037/g4667∗/g3090/g3004/g3042/g2878/g3030/g3008/g3031 /g4666/g3037/g4667∗/g3090/g3008/g3031 /g4667/g4673/g3036\n/g3037/g2880/g2868 /g4673 with c x(j) \nand λx the concentration of the specie x in the layers j (on top of the layer i) and the 5 \n photoelectron inelastic mean free path (IMFP) respe ctively. The total XPS signal of specie x is \nthen given by /g1827/g3051=/uni2211 /g1829/g3051/g3036/g1827/g3051/g3036 /g3015\n/g3036/g2880/g2868 with N the alloy thickness and R is given by /g3002/g3256/g3279 \n/g3002/g3252/g3290 . We have used \na genetic algorithm in order to determine the profi le which gives the best match between the \ncalculated dependence of R on photon energy and the experimental values (figure 6a). \nConsidering the limited inelastic mean free path of photoelectrons in CoGd alloys (table 1), we \nhave set the composition to the nominal value as gi ven by VSM measurements (figure 1) for \nthicknesses above 5 nm. \nThe laser induced ultrafast Gd 4f spin dynamics was recorded by using TR-DRAS \n[BEA13, SIL17] at the TEMPO Beamline of synchrotron SOLEIL [POL10] by using both the \nhybrid (figure 7) and low-alpha (figure 8) filling mode [SIL17]. This technique offers surface \nsensitivity, element specificity, the sensitivity t o 4f magnetic order and time-resolutions of 60 \nps or 12 ps in the hybrid and low-alpha filling mod e, respectively. In our experiments, the \nenergy of the circularly polarized X-ray pulses mat ches the Gd M 5 absorption edge \n(E hν=1192.7eV) while the Auger photoelectrons (E KE =1184.5eV) are collected by a Scienta \n2002 photoelectron analyzer equipped with a delay l ine [BER11]. The kinetic energies of Auger \nphotoelectrons allow separating them from the photo electrons generated by laser absorption \n[SIR14]. The intensity of the generated Auger photo electrons (in e-/sec) is proportional to the \nX-ray absorption in the Co xGd 100-x layers. Since the X-ray absorption of circularly p olarized X-\nray in magnetic materials depends on the magnetizat ion, the Auger photoelectron yield depends \non the magnetization, as depicted in figures 7(a,b) and figures 8(a,b). The difference between \nthe Auger photoelectrons collected for two opposite directions of the magnetic field (H+ and \nH-) is proportional to the magnetization [SIL17]. T he variation in the amplitude of the signal \nproportional to the magnetization is indicated by t he blue M arrows in the figures. The TR-\nDRAS experiments have consisted in recording the Au ger photoelectrons distribution for both \nmagnetic field helicities as a function of the dela y between the IR pump and the X-ray probe \npulses by using either the hybrid (time resolution ~60 ps) or the low alpha (time resolution \n~12ps) modes. In figures 7(a,b) and 8(a, b), we dis play the Auger photoelectron yield for two \nmagnetic field helicities at negative delay (a) and t=100 ps after laser excitation (b) for hybrid \nand low-alpha modes, respectively. Then, the photoe lectron Auger yields are integrated over \nthe investigated kinetic energy range. The differen ce of the integrated signal between both \nmagnetic field helicities as a function of the pump -probe delay (figure 7c and 8c) is labelled \nTR-DRAS in the following. The laser frequency was s et to 141 kHz, which is six times lower \nthan the frequency of the isolated electron bunch i n the SOLEIL filling pattern. The thick 6 \n Ta/Cu/Ta buffer layer beneath the Co xGd 100-x alloys allows to enhance the heat dissipation \nduring the pump-probe experiments, ensuring moderat e temperature elevation by DC-heating. \nThe laser beam impinges the sample at an angle of 6 7° in respect with the sample’s normal \n(figure 2). For each delay step, a pulsed magnetic field of 300 ms duration, 200 Oe amplitude \nis applied along the magnetic easy axis to saturate the alloys. The acquisition is then performed \nat remanence (H ext = 0 Oe), which prevents perturbation of the photoe lectron trajectories \ntowards the analyzer. The detection scheme allows c ollecting separately the photoelectron \npulses generated by every isolated X-ray pulses, wh ich means that between two successive laser \nexcitations, 1 X-ray pulse probes the excited magne tic state of the alloy while the next 5 pulses \nprobe the non-excited magnetic state of the alloy. Thus, we have direct evidences that the \nsamples are not damaged during the acquisition and that the magnetization of the alloy is \nrestored to its equilibrium state between two succe ssive excitations even in absence of external \nmagnetic field. It also allows normalizing the photoelectron Auger yield coming from the \nexcited magnetic state by the photoelectron Auger y ield coming from the non-excited magnetic \nstate to increase significantly the signal to noise ratio [SIL17]. The time-resolved experiments \nwere carried out for the Co 65 Gd 35 and Co 80 Gd 20 alloys (figures 7 and 8). We have not fully \nremoved the Al protective layer during the time-con suming pump-probe experiments to avoid \noxidization of the top layers [BER17]. The cryostat temperatures were set to 80 and 200 K \nwhile the laser fluences were 2.1 and 2.8 mJ/cm² re spectively. The higher temperature for the \nCo 80 Gd 20 alloy was needed to reach a coercive field below 20 0 Oe, while the larger laser fluence \nwas selected to obtain similar demagnetization ampl itudes. We have restricted the laser fluence \nto 2.8 mJ/cm² because we observed traces of degradation at the su rface of the alloys for laser \nfluence above 3 mJ/cm². Furthermore, we have restri cted our investigations to moderate \nexcitation (~50 % demagnetization) for direct compa rison with previous bulk-sensitive XMCD \nexperiments performed in pure Gd [WIE11, ESC14] and CoGd alloys [FER23]. As a \nconsequence, for the Co 65 Gd 35 (Co 80 Gd 20 ) alloy, the equilibrium temperature was below \n(above) the temperature of magnetic compensation [F ER17]. The laser induced dynamics of \nGd 4f spins in the Co 65 Gd 35 alloy was also recorded on a broader delay range by using the \nhybrid mode (figure 7). Such measurements allow to determine accurately the characteristic \nrecovery times in order to lower the uncertainties in fitting the TR-DRAS data obtained in the \nlow-alpha mode. For the Co 80 Gd 20 alloy, the delay range investigated in the low-alp ha operation \nmode was sufficient to estimate the characteristic recovery times. \nExperimental results and discussion : 7 \n The net magnetization of the alloys as a function of temperature measured by PPMS-\nVSM magnetometry and XMCD spectroscopy are plotted in figure 1 and figure 3 respectively. \nWe observe that the magnetization of the Co 65 Gd 35 alloy goes to zero at T ~ 370 K (figure 1), \nas expected from tabulated data [TAO74, HAN89]. We also observe that the magnetization of \nthe Co 80 Gd 20 alloy crosses zero at T = 170 K (figure 1), which i s the temperature of magnetic \ncompensation (T comp ). This value for T comp is also consistent with the nominal composition as \nattested by tabulated data [HAN89]. The opposite si gns of the XMCD signal at the Gd M 5 edges \nfor the Co 80 Gd 20 alloy (figure 3b) and the Co 65 Gd 35 alloy (figure 3d) confirm that the XMCD \nspectra were acquired above (below) the temperature of magnetic compensation for the \nCo 80 Gd 20 (Co 65 Gd 35 ) alloy. The characterization of the magnetic propert ies by mean of PPMS-\nVSM magnetometry and XMCD spectroscopy show that th e actual composition of the alloy in \nthe bulk matches the nominal composition. \nAfter ensuring that the alloys have the correct com position in the bulk, we focused on \nthe electronic and structural properties at the sur face of the alloy by means of XPS (figure 4). \nIn figure 5a and 5b, we observe that the height of the Co 3p peak increases in respect with that \nof the Gd 4f peak when the photon energy increases for both samples. It demonstrates that the \nsurface of the alloys is less concentrated in Co th an the bulk. Our model gives a qualitative \nagreement with the experimental variation of R (fig ure 6a). The profile which corresponds to \nthe calculated variation of R with photon energy is depicted in figure 6b. In the case of Co 80 Gd 20 \nalloy, the simulated profile shows a segregated Gd layer on top of an almost homogeneous \nsample with almost the nominal composition as previ ously reported for a comparable alloy \ncomposition [BER17]. However, the simulated profile of the Co 65 Gd 35 alloy shows a \npronounced Gd compositional gradient at the surface of the alloy and thus, a sizable increase \nof the Gd contents over almost 2 nm in respect with the nominal composition. A more accurate \ndetermination of the profile given for layers deepe r than 2 nm should be confirmed by other \nexperimental techniques with higher probing depth. Our observations based on XPS \ndemonstrate that the alloys are non-homogeneous and that these non-homogeneity’s \ndiscrepancies depend on the alloy composition. \nThe TR-DRAS measurements show a quenching of the Gd 4f magnetic order which is \nfollowed by a recovery for both alloys (Figure 7c, 8c). However, our data show different \ndynamical responses to laser excitation for the Gd sublattices in both alloys. Indeed, we observe \nthat the maximum demagnetization amplitudes are rea ched at delays d max = 28 ps and d max = \n100 ps for the Co 80 Gd 20 and the Co 65 Gd 35 alloys respectively (figure 8). The data were adjus ted 8 \n by exponential functions convoluted with a Gaussian function to extract the characteristic \ndemagnetization ( τM) and recovery ( τR) times (solid lines in figures 7c and 8c) [BOE10, LOP12, \nBER14]. We have estimated τM = 12 ps and τR = 55±10 ps for the Co 80 Gd 20 alloy and τM = \n37±10 ps and τR = 400±60 ps for the Co 65 Gd 35 alloy. For the Co 80 Gd 20 alloy, we have set a lower \nlimit of 12 ps for τM during the fit processing which is given by the ex perimental time resolution. \nFor the Co 65 Gd 35 alloy, the recovery time ( τR = 400±60 ps) was extracted from the data acquired \nin the hybrid mode (figure 7c). \nThe faster recovery for the Co 80 Gd 20 alloy ( τR = 55±10 ps) measured at T = 200K \ncompared to the Co 65 Gd 35 alloy ( τR = 400±60 ps) measured at T = 80K is probably partl y caused \nby the temperature dependent thermal conductivity i n RE-TM alloys as shown by Hopkins et \nal. in FeCoGd alloys [HOP12]. Such temperature depe ndent recovery for Gd 4f spins in CoGd \nalloys has been reported recently [FER23]. Furtherm ore, the article published by Ferté et al. \nallows explaining that the characteristic demagneti zation time τM < 12 ps for the Co 80 Gd 20 alloy \nis shorter than that for the Co 65 Gd 35 alloy ( τM = 37±10 ps), because of the different Gd \ncomposition of the alloys. Indeed, in CoGd alloys, the indirect Gd 4f spin-lattice coupling is \nenhanced in respect with pure Gd layers by the Gd 4 f – Gd 5d intra-atomic exchange coupling, \nthe Co 3d – Gd 5d inter-atomic exchange coupling an d the Co 3d spin-lattice coupling [REN19, \nHEN19, FER23]. By reducing the Co concentration at the surface of the alloy, the indirect Gd \n4f spin-lattice coupling is also reduced. However, the laser induced dynamics of Gd 4f spins on \ntop of the Co 65 Gd 35 alloy shows similarities with that reported by Wiet struk et al. for pure Gd \nlayer in low-alpha mode at BESSY [WIE11]. Indeed, a lthough the time-resolution of our TR-\nDRAS experiment and the statistic do not allow for resolving the two-step demagnetization, it \nis worth noticing that the demagnetization times we extracted from our experimental data ( τM \n= 37±10 ps) is similar to the characteristic times of the “slow” demagnetization process of pure \nGd [WIE11] rather than that reported by means of bu lk-sensitive TR-XMCD in Co 72 Gd 28 alloy \n(τM ~ 4 ps [FER23]). This observation is thus qualitat ively consistent with the rich Gd content \nat the surface of the Co 65 Gd 35 alloys, which induces disparate Gd 4f spin dynamic s compared \nto that expected within the bulk. It is worth notic ing that Graves et al. have shown that lateral \ncomposition gradient also influence the spin dynami cs in RE-TM alloys [GRA13]. However, \nthe amplitude in composition non-homogeneities they have reported are much lower than that \nwe estimate at the surface of our layers. In the ca se of the Co 80 Gd 20 alloy, any discussion \nregarding the characteristic demagnetization time f or the Gd 4f sublattice and its comparison \nwith previous bulk-sensitive measurements [LOP13, R AD15] are hampered by the time-9 \n resolution of our experiment. Our works show that a n accurate description of the Gd 4f spin \ndynamics in ferrimagnetic alloys requires particula r attention to their structural properties. Our \nwork also calls for further in-depth resolved exper iments in RE-TM alloys with nanometer \nspatial resolution combined with sub-picosecond tim e resolutions either at large scale facilities \n[JAL17] or by using table-top X-ray sources [HEN22] . \n \nConclusions: \nIn this work, we propose an experimental investigat ion on the structural and static \nmagnetic properties as well as on the laser induced ultrafast Gd 4f spin dynamics at the surface \nof Co xGd 100-x ferrimagnetic alloys. We have shown that the surfa ce of the CoGd alloys is Gd-\nrich compared to the bulk and these non-homogeneiti es are composition dependent. Indeed, the \nGd migration at the surface is more pronounced in a lloys that display a larger Gd content in the \nbulk. We have used surface sensitive time-resolved dichroic resonant Auger spectroscopy to \nshow that the Gd 4f spin dynamics is much slower at the surface of Co 65 Gd 35 alloy compared \nto that previously reported in the bulk of comparab le alloys by means of time-resolved XMCD \nmeasurements. This discrepancy can be explained by the larger Gd composition on the top of \nthe alloys which results in a reduced indirect Gd 4 f spin-lattice coupling. However, our \npreliminary work gives a partial view of the Gd 4f spin dynamics because of the limited time-\nresolution and calls for further in-depth and time- resolved experiments. Among the point to be \nclarified, we hint that the exchange coupling betwe en the Gd-rich surface and the bulk of the \nalloy may also play a role on the Gd 4f spin dynami cs by modifying the effective Curie \ntemperature (T Curie ). T Curie is known to be a key parameter to determine the ch aracteristic \ndemagnetization times in the framework of the micro scopic 3 temperatures model [KOO10]. \nThese issues have to be addressed for future techno logical applications which require a fine \ntuning of the characteristic demagnetization times [REM20] or a downsizing of the \nferrimagnetic alloy layers [LAL17, SEN21]. \nAcknowledgements: \n \nWe acknowledge SOLEIL for provision of synchrotron radiation facilities in using the beamline \nTEMPO. This project has received fundings from the European Union’s Horizon 2020 research and \ninnovation program under the Marie Skłodowska-Curie grant agreement number 847471, the French \nnational agency for research ANR-20-CE42-0012-01 an d from the Région Grand Est. The competence 10 \n center Magnetism and Cryogenics, Institut Jean Lamo ur, is acknowledged for magnetometry \nmeasurements. \n \nReferences: \n \n[OST12] Ostler et al. Nature Commun. 3, 666 (2012) \n[RAD11] Radu et al, Nature. 472, 205 (2011) \n[LOP13] V. López-Flores, N. Bergeard, V. Halté, C. Stamm, N. Pontius, M. Hehn, E. Otero, E. \nBeaurepaire, and C. Boeglin, Phys. Rev. B 87, 21441 2 (2013) \n[GRA13] Graves et al. Nature Materials. 12, 293 (2013) \n[BER14] Bergeard et al. Nature Communications 5, 34 66 (2014) \n[RAD15] I. Radu, C. Stamm, A. Eschenlohr, F. Radu, R. Abrudan, K. Vahaplar, T. Kachel, N. \nPontius, R. Mitzner, K. Holldack, and A. Föhlisch, SPIN 5, 1550004 (2015) \n[HIG16] Higley et al. Rev. of Sci. Instrum. 87, 033110 ( 2016 ) \n[HEN19] Hennecke et al. Phys. Rev. Lett. 122, 157202 (2019) \n[IIH18] Iihama et al. Advanced Materials 30, 1804004 (2018) \n[FER17] Ferté et al. Phys. Rev. B. 96, 134303 (2017 ) \n[CHE19] Chen et al. New J. Phys. 21, 123007 (2019) \n[REN20] Ren et al. arXiv:2012.14620v1 (2020) \n[FER21] Ferté et al. J. of Magn. and Magnetic Mat. 530, 167 883 (2021) \n[FER23] Ferté et al. Eur. Phys. J. ST 232, 2213 (2023) \n[HEN22] Hennecke et al. Phys. Rev. Research 4, L022 062 (2022) \n[ATX14] U. Atxitia, J. Barker, R. W. Chantrell, and O. Chubykalo-Fesenko, Phys. Rev. B 89 , \n224421 (2014). \n[HEB16] Hebler et al. Front. Mater. 3, 8 (2016) \n[HAL18] Haltz et al. Phys. Rev. Mat. 2, 104410 (2018) \n[INY23] Inyang et al. Appl. Phys. Lett. 123, 122403 (2023) \n[SHE81] Shen et al. Jap. J. Appl. Phys. 20, L757 (1 981) \n[BER17] N. Bergeard, A. Mougin, M. Izquierdo, E. Fo nda, and F. Sirotti, Phys. Rev. B 96, \n064418 (2017) \n[SEN21] Seng et al. Adv. Funct. Mater. 31, 2102307 (2021) \n[BEA13] Beaulieu et al. Journal of Elec. Spec. 189, 40 (201 3) \n[SIL17] Silly et al. Journal of synchr. Rad. 24, 886 (2017) \n[NAK99] Nakajima et al. Phys. Rev. B 59, 6421 (1999 ) 11 \n [POL10] Polack et al. AIP Conference Proceedings, 1 234, 185 (2010) \n[TAY76] Taylor et al. J. Appl. Phys. 47, 4666 (1976 ) \n[JAB11] Jablonski et al. Auger Electron Spectroscop y, Version 1.0, National Institute of \nStandards and Technology, Gaithersburg, Maryland (2 011). \n[YEH85] Yeh et al. At. Data Nucl. Data Tables 32, 1 (1985) \n[PAC23] Pacé. « Génération de courants d’électrons polarisés en spi n par désaimantation \nultra-rapide : processus fondamentaux et applicatio ns pour l’électronique THz », PhD thesis \nUniversity of Strasbourg (2023) \n[BER11] Bergeard et al. J. of Synchrotron Rad. vol 18, 245 (2011) \n[SIR14] Sirotti et al. PRB 90, 035401 (2014) \n[WIE11] M. Wietstruk, A. Melnikov, C. Stamm, T. Kac hel, N. Pontius, M. Sultan, C. Gahl, M. \nWeinelt, H. A. Dürr, and U. Bovensiepen, Phys. Rev. Lett. 106, 127401 (2011) \n[ESC14] A. Eschenlohr, M. Sultan, A. Melnikov, N. B ergeard, J. Wieczorek, T. Kachel, C. \nStamm, and U. Bovensiepen, Phys. Rev. B 89, 214423 (2014) \n[TAO74] Tao et al. AIP Conf. Proc. 18, 641 (1974) \n[HAN89] Hansen et al. J. Appl. Phys. 66, 756 (1989) \n[BOE10] Boeglin et al. Nature 465, 458 (2010) \n[LOP12] Lopez-Flores et al., Phys. Rev. B. 86, 0144 24 (2012) \n[HOP12] Hopkins et al. JAP 111, 103533 (2012) \n[JAL17] Jal et al. Phys. Rev. B 95, 184422 (2017) \n[KOO10] Koopmans et al. Nature Materials 9, 259 (2010) \n[REM20] Rémy et al. Adv. Sci. 2001996 (2020) \n[LAL17] Lalieu et al. Phys. Rev. B 96, 220411(R) (2 017) \n \nTable 1: Mean free paths of the different core-level photoel ectrons Co3d, Co3p, Gd4f and Gd \n4d as a function of the incident photon energy as g iven by [JAB11]. \n \n \nPhoton energy \n(eV) Mean free path λ (Å) \nλCo 3d λCo 3p λGd 4f λGd 4d \n400 6.6 6.2 9.8 7.3 12 \n 700 8.5 8.2 14.7 12.5 \n1000 10 9.7 19.4 17.3 \n \nFigures: \n \n \nFigure 1: Magnetization as a function of temperatur e measured by mean of VSM magnetometry for the Co 80 Gd 20 (black \nfilled squares) and Co 65 Gd 35 (red empty circles) alloys. The values are normali zed by the magnetization at T = 80K. \n \n \nFigure 2: Geometry of the experiment where the exte rnal magnetic field can be applied in the plane of the films. Incidence \nangles of the X rays and IR pump are defined by α X (=44°) and α IR (=67°) in respect with the surface normal. \n \n13 \n \nFigure 3: X-ray absorption spectra for two opposite helicities of the external magnetic field (black a nd red curves) at the Gd \nM4,5 absorption edges for the Co 80 Gd 20 alloy at T = 220K (a) and for the Co 65 Gd 35 alloy at T = 80 K (c). (b) The XMCD spectra \nat the Gd M 4,5 absorption edges for the Co 80 Gd 20 alloy at T = 220K and T = 300K. (d) The XMCD spect ra at the Gd M 4,5 \nabsorption edges for the Co 65 Gd 35 alloy at T = 80K, T = 200K and T = 300K. \n \n \nFigure 4: X-ray photoelectron spectra taken at 700 eV photon energy showing the Co 3p , 3d, Gd 4f ,4d and the Al 2p core-\nlevels for the alloys (a) Co 80 Gd 20 and (b) Co 65 Gd 35 . \n \n \nFigure 5: X-ray Photoelectron spectra of the Co 3p and Gd 4f core-levels for (a) Co 80 Gd 20 and (b) Co 65 Gd 35 at the photon \nenergies of 400 eV (black spectra), 700 eV (red spe ctra) and 1000 eV (green spectra). The spectra have been subtracted by a \n14 \n Shirley function and normalized by the photoionizat ion cross section [JAB11]. All spectra have been no rmalized to 1 at the \nGd4f. \n \n \nFigure 6: (a) Comparison between the experimental ( symbols) and the calculated (solid lines) energy de pendence of the ratio \nR between the Gd 4f and Co 3p core levels. (b) Simu lated Gd composition profile at the surface of the Co 80 Gd 20 (black line) \nand the Co 65 Gd 35 (red line) alloys \n \nFigure 7: Auger electron yield for two magnetic fie ld helicities at negative (a) and positive (b) dela ys for the Co 65 Gd 35 alloy \nmeasured by using the hybrid filling mode (time res olution 60 ps). (c) Normalized magnetic circular di chroism as a function \nof the pump-probe delay for the Co 65 Gd 35 alloy (red filled circle) at T = 80 K. The solid li nes are exponential fit as described in \ntext. \n15 \n \nFigure 8: Auger electron yield for two magnetic fie ld helicities at negative (a) and positive (b) dela ys for the Co 65 Gd 35 alloy \nmeasured by using the low-α filling mode (time-reso lution 12 ps). (c) Normalized magnetic circular dic hroism as a function of \nthe pump-probe delay for the Co 65 Gd 35 (red filled circle) at T = 80 K and the Co 80 Gd 20 at T = 200 K (black empty squares) \nalloys. The solid lines are exponential fit as desc ribed in text. \n \n \n" }, { "title": "1409.8430v2.Predicting_a_Ferrimagnetic_Phase_of_Zn2FeOsO6_with_Strong_Magnetoelectric_Coupling.pdf", "content": " \n1 \n Predicting a Ferrimagnetic Phase of Zn 2FeOsO 6 with Strong Magnetoelectric Coupling \nP. S. Wang ,1 W. Ren,2 L. Bellaiche,3 and H. J. Xiang*1 \n \n1Key Laboratory of Computational Physical Sciences (Ministry of Education), State Key \nLaboratory of Surface Physics, Collaborative Innovation Center of Advanced Microstructures , \nand Department of Physics, Fudan University, Shanghai 200433, P. R. China \n2Depa rtment of Physics, Shanghai University, 99 Shangda Road, Shanghai 200444, P. R. China \n3Physics Department and Institute for Nanoscience and Engineering, University of Arkansas, \nFayette ville, Arkansas 72701, USA \nAbstract \nMultiferroic materials, in which ferroelectric and magnetic ordering coexist, are of fundamental \ninterest for the development of novel memory devices that allow for electrical writing and non -\ndestructive magnetic readout operation. The great challenge is to create roo m temperature \nmultife rroic materials with strongly coupled ferroelectric and ferromagnetic (or ferrimagnetic) \nordering s. BiFeO 3 has been the most heavily investigated single -phase multiferroic to date due to \nthe coexistence of its magnetic order and ferroe lectric order at room temperature. However, there \nis no net magnetic moment in the cycloidal ( antiferromagnetic -like) magnetic state of bulk \nBiFeO 3, which severely limits its realistic applications in electric field controlled spintronic \ndevices. Here, we predict that double perovskite Zn 2FeOsO 6 is a new multiferroic with properties \nsuperior to BiFeO 3. First, there are strong ferroelectricity and strong ferrimagnetism at room \ntemperature in Zn 2FeOsO 6. Second, the easy -plane of the spontan eous magnetization can be \nswitched by an external electric field, evidencing the strong magnetoelectric coupling existing in \n2 \n this system. Our results suggest that ferrimagnetic 3d -5d double perovskite may therefore be \nused to achieve voltage control of mag netism in future spintronic devices. \nPACS: 75.85.+t, 75.50.Gg, 71.15.Mb, 71.15.Rf \n \n \n The highly efficient control of magnetism by an electric field in a solid may widen the \nbottle -neck of the state -of-the-art spin -electronics (spintronics) technology, such as magnetic \nstorage and magnetic random -access memory. Multiferroics [1-7], which show simultaneous \nferroelectric and magnetic ordering s, provide an ideal platform for the electric field control of \nmagnetism because of the coupling between their dual order parameters. For realistic \napplications, one need s to design/discover room temperature multiferroic materials with strong \ncoupled ferroelectric and ferromagnetic (or ferrimagnetic) ordering. \n Perovskite -structure bismuth ferrite (BiFeO 3) is currently the most studied room \ntemperature single -phase multiferroic, mostly because its large polarization and high \nferroelectric Curie temperature (~820 ° C) make it appealing for applications in ferroelectric non -\nvolatile memories and high temperature electronics. Bulk BiFeO 3 is an antiferromagnet with \nNé el temperature T N ≈ 643 K [8]. The Fe magnetic moments order almost in a checkerboard G -\ntype manner with a cycloidal spiral spin structure in which the antiferromagnetic (AFM) axis \nrotates through the crystal with an incommensurate long -wavelength period [9]. This spiral spin \nstructure leads to a cancellation of any macroscopic magnetization. The magnetic properties of \nBiFeO 3 thin films were found to be markedly different from those of the bulk: The spiral spin \nstructure seems to be suppressed and a weak magnetization appears [10]. Nevertheless, the \nmagnetization is too small for many applications [11]. In addition, an interesting low -field \n3 \n magnetoelectric (ME) effect at room temperature was discovered in Z -type hexaferrite \nSr3Co2F24O41 by Kitagawa et al. [12]. Unfortunately, the electric polarization (about 20 µC/m2) \ninduced by the spin order is too low [13]. \nIn searching for new multiferroic compounds, the double perovskite system was proposed \nas a promising candidate [14, 15]. The double perovskite structure A 2BB′O6 is derived from the \nABO 3 perovskite structure. The two cations B and B ′ occupy the octahedra l B sites of perovskite \nwith the rock salt ordering. Double perovskite Bi 2NiMnO 6 was successfully synthesized under \nhigh-pressure, which displays the multiferroic behavior with a high ferroelectric transition \ntemperature (485 K) but a low ferromagnetic transition temperature (14 0 K) [14]. Polar LiNbO 3 \n(LN) -type Mn 2FeMO 6 (M=Nb, Ta) compounds were prepared at 1573 K under 7 GPa [16]. \nUnfortunately, the magnetic ground state of Mn 2FeMO 6 is AFM with a rather low Né el \ntemperature (around 80 K). Very recently, LN -type polar magnetic Zn 2FeTaO 6 was obtained via \nhigh pressure and temperature synthesis [17]. The AFM magnetic transition temperature (T N∼22 \nK) for Zn2FeTaO 6 is also low. In a pioneering work, Ležaić and Spaldin proposed to design \nmultiferroics based on 3d -5d ordered double perovskites [18]. They found that Bi 2NiReO 6 and \nBi2MnReO 6 are insulating and exhibit a robust ferrimagnetism that persists above room \ntemperature. Although coherent heteroepitaxy strain may stabilize the R3 ferroelectric ( FE) state, \nfree-standing bulk of Bi2NiReO 6 and Bi 2MnReO 6 unfortunately take the non -polar P21/n \nstructure as the ground state. The magnetic properties of non -polar double perovskites \nCa2FeOsO 6 [19] and Sr 2FeOsO 6 [20] were also theoretically investigated. Recently, Zhao et al. \npredicted that double perovskite superlattices R 2NiMnO 6/La 2NiMnO 6 (R is a rare -earth ion) \nexhibit an electrical polarization and strong ferromagnetic order near room temperature [21]. \n4 \n However, the ME coupling in these superlattices appears to be weak and the polariz ation to be \nsmall . \nIn this work, we predict that double perovskite Zn 2FeOsO 6 takes the FE LN -type structure \nas the ground state through a global structure searching. Similar to Bi 2NiReO 6 and Bi 2MnReO 6, \nZn2FeOsO 6 exhibit a strong ferrimagnetism at room temperature. Importantly, there is a rather \nstrong magnetic anisotropy with the easy -plane of magnetization perpendicular to the FE \npolarization due to the presence of the significant 3d -5d Dzyaloshinskii -Moriya (DM) interaction . \nThis suggests that the swi tching between the 71 °or 109°FE domains by the electric field will \ncause the rotation of the magnetic easy -plane. Our work therefore indicates that Zn 2FeOsO 6 may \nbe a material of choice for realizing voltage control of magnetism at room temperature. \nIt is well-known that there are several lattice instabilities including ferroelectric distortion s \nand oxygen octahedron rotation s in perovskite materials . We now examine how double -\nperovskite Zn 2FeOsO 6 distort s to lower the total energy. For this purpose, we per form a global \nsearch for the lowest energy structure based on the genetic algorithm (GA) specially designed for \nfinding the optimal structural distortion [22]. We repeat the simulations three times. All three \nsimulations consistently show that the polar rhombohedral structure with the R3 space group \n[shown in Fig. 1( b) and ( c)] has the lowest energy for Zn 2FeOsO 6. Similar to Zn 2FeTaO 6, \nZn2FeOsO 6 with the R3 structure is based on the R3c LN -type structure. Previous experiments \nshowed that double perovskite structure A 2BB′O6 may adopt other structures, such as the P21/n \n[23], \n3R[24], and C2 structure s [25]. Our density functional theory ( DFT ) calculations show \nthat the R3 phase of Zn2FeOsO 6 has a lower energy than the P21/n, \n3R , and C2 structures by \n0.22, 0.09, 0.45 eV/f.u., respectively. This strongly suggest s that double perovskite Zn 2FeOsO 6 \n5 \n adopts the R3 structure as its ground state. This can be understood by using the tolerance factor \ndefined for the LN -type ABO 3 system. It was shown that when the tolerance factor \n(\n2( )AO\nR\nBOrrt\nrr\n , where \nAr , \nBr and \nOr are the ionic radii of the A -site ion, B -site ion and O ion ) \nis smaller than 1, the polar \n3Rc structure is more stable than the \n3Rc ABO 3 structure due to the \nA-site instability [26]. In the case of Zn 2FeOsO 6, the average tolerance factor ( 0.75) is smaller \nthan 1, which suggests that the \n3R structure is more stable than the non -polar \n3R structure. \nNote that phonon calculation s shows that the \n3R state of Zn2FeOsO 6 is dynamically stable [27]. \nAdditional tests indicate that the Fe and Os ions tend to order in a rock salt manner (i.e., double \nperovskite configuration ) to lower the Coulomb interaction energy [27,28]. \nOur electronic structure calculation shows that the R3 phase of Zn2FeOsO 6 in the \nferrimagnetic state is insulating [27]. The density of states plot shows that the Fe majority 3d \nstates are almost fully occupied, while the minority states are almost empty. This suggests that \nthe Fe ion takes the high -spin Fe3+ (d5) valence state. It is also clear that the Os ion takes the \nhigh-spin Os5+ (d3) valence state, which contra sts with the case of Ba2NaOsO 6 where the Os \natom takes a 5d1 valence electron configuration [29]. This is also consistent with the total \nmagnetic moment of 2 µB/f.u. for the ferrimagnetic state from the collinear spin -polarized \ncalculation . Through the four -state mapping approach which is able to deal with spin in teractions \nbetween two different atomic types [30], we compute the symmetric exchange parameters to find \nthat the magnetic ground state of R3 Zn 2FeOsO 6 is indeed ferrimagnetic. The Fe -Os \nsuperexchange interactions mediated by the corner -sharing O ions [J 1 and J 2, see Fig. 1( c)] are \nstrongly AFM (J 1 = 31.68 meV , J2 = 29.62 meV ). Here, the spin interaction parameters are \n6 \n effective by setting the spin values of Fe3+ and Os5+ to 1. Our results seem to be in contradiction \nwith the Goodenough -Kanamori rule which predicts a ferromagnetic interaction between a d5 ion \nand a d3 ion since the virtual electron transfer from a half -filled σ -bond e orbital on the d5 ion to \nan empty e orbital on the d3 ion dominates the antiferromag netic π -bonding t -electron transfer \n[31]. This discrepancy is because the \ndirections. In rhombohedral Zn2FeOsO 6, the spontaneous electric polarization is directed along \none of the eight <111> axes of the perovskite structure. Thus, in the sample of double -perovskite \nZn2FeOsO 6, there might occur eight different FE domains [see Fig. 4( b)]. Our above calculations \nshow that the easy -plane of mag netization is always perpendicular to the direction of the electric \npolarizatio n. Although a 180° switching of the ferroelectric polarization should not affect the \nmagnetic state, a 71° or 109° switch of the FE domains by the electric field will change the \norientation of the easy -plane of magnetization, as shown schematically in Fig. 4( c). This could \nbe a promising route to manipulate the orientation of the ferrimagnetism by an electric field. A \nsimilar ME coupling mechanism in BiFeO 3 thin films has been de monstrated experimentally by \nZhao et al. , who showed that the AFM plane can be switched by an electric field [39]. Note that \nmagnetoelectric effects can be classified in to two different types : one for which changing the \nmagnitude of the polarization affects the magnitude of the magnetization (energy of the form\n22PM\n) and one for which changing the direction of \nP\n changes the direction of \nM\n (energy of \nthe form \nPM\n ). In Zn 2FeOsO 6, the first type of ME effect is weak, while the second type of ME \neffect is strong. \nWe now compare Zn2FeOsO 6 with the classic multiferroic BiFeO 3. First, they adopt similar \nrhombohedral structures. Second, both compounds have high electric al polarizations. Third, both \n10 \n compounds are room temperature multiferroics. Fourth, the ME coupling mechanism is rather \nsimilar in that the magnetic easy -plane can be manipulated by electric field. However, the \nmagnetic ground state of R3 Zn2FeOsO 6 is dramatically different from BiFeO 3. Zn2FeOsO 6 has a \nferrimagnetic ground state, while BiFeO 3 is AFM. And the magnetic anisotropy in Zn2FeOsO 6 is \nstronger than that (0.2 meV when U(Fe) = 5 eV [40]) in BiFeO 3 because of the strong SOC \neffect of the 5d Os ion. These desirable properties make Zn2FeOsO 6 suitable for realizing \nelectric -field control of magnetism at room temperature. \nThe reason why we practically propose d Zn2FeOsO 6 as a possible multiferroic is two -fold. \nFirst, polar Zn 2FeTaO 6 has already been synthesized under high -pressure, as mentioned above. \nSecond, it was experimentally showed that Ca 2FeOsO 6 crystallizes into an ordered double -\nperovskite structure with a space gro up of P21/n under high -pressure and high -temperature, and \nCa2FeOsO 6 presents a long -range ferrimagnetic transition above room temperature (T c ∼320 K) \n[24]. Therefore, we expect that Zn 2FeOsO 6 is synthesizable and would most likely exhibit both \nferroelectricity and ferrimagnetism at room temperature. \nWork at Fudan was supported by NSFC, FANEDD, NCET -10-0351, Research Program of \nShanghai Municipality and MOE, the Special Funds for Major State Basic Research, Program \nfor Professor of Special Appoin tment (Eastern Scholar), and Fok Ying Tung Education \nFoundation. L.B. thanks the Department of Energy, Office of Basic Energy Sciences, under \ncontract ER -46612 . \ne-mail: hxiang@fudan.edu.cn \nReferences \n[1] S.-W. Cheong and M. Mostovoy, Nature Mater. 6, 13 (2007). \n11 \n [2] R. Ramesh and N. Spaldin , Nature Mater. 6, 21 (2007). \n[3] S. Picozzi and C. Ederer, J. Phys. : Condens. Matter 21, 303201 (2009) . \n[4] K. F. Wang, J.-M. Liu, and Z. Ren, Adv. Phys. 58, 321 (2009) . \n[5] J. van den Brink and D. Khomskii, J. Phys .: Condens . Matter 20, 434217 (2008). \n[6] Y . Tokura and S. Seki, Adv. Mater. 22, 1554 (2010). \n[7] J. Ma, J. M. Hu, Z. Li, and C. W. Nan, Adv. Mater. 23, 1062 (2011). \n[8] S. V. Kiselev, R. P. Ozerov, and G. S. Zhdanov, Sov. Phys. Dokl. 7, 742 (1963); D. \nRahmedov, D. Wang, J. Í ñ iguez, and L. Bellaiche, Phys. Rev. Lett. 109, 037207 (2012); I. C. \nInfante, S. Lisenkov, B. Dupé , M. Bibes, S. Fusil, E. Jacquet, G. Geneste, S. Petit, A. Courtial, J. \nJuraszek, L. Bellaiche, A. Barthé lé my, and B. Dkhil, Phys. Rev. Lett. 105, 057601 (2010). \n[9] I. Sosnowska, T. Peterlin -Neumaier, and E. Streichele, J. Phys. C 15, 4835 (1982). \n[10] J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D. Viehland, V. \nVaithyanathan, D. G. Schlom, U. V . Waghmare, N. A. Spaldin , K. M. Rabe, M. Wuttig, and R. \nRamesh, Science 299, 1719 (2003). \n[11] W. Eerenstein , F. D. Morrison , J. Dho, M. G. Blamire , J. F. Scott , and N. D. Mathur , \nScience 307, 1203 (2005). \n[12] Y. Kitagawa , Y. Hiraoka , T. Honda , T. Ishikura , H. Nakamura , and T. Kimura , Nature Mater. \n9, 797 (2010). \n[13] H. J. Xiang, E. J. Kan, Y. Zhang, M.-H. Whangbo, and X. G. Gong, Phys. Rev. Lett. 107, \n157202 (2011); H. J. Xiang, P. S. Wang, M.-H. Whangbo, and X. G. Gong, Phys. Rev. B 88, \n054404 (2013). \n[14] M. Azuma, K. Takata, T. Saito, S. Ishiwata, Y. Shimakawa, and M. Takano, J. Am. Chem. \nSoc. 127, 8889 (2005 ). \n12 \n [15] S. Kumar, G. Giovannetti, J. van den Brink, and S. Picozzi, Phys. Rev. B 82, 134429 (2010). \n[16] M.-R. Li, D. Walker, M. Retuerto, T. Sarkar, J. Hadermann, P. W. Stephens , M. Croft , A. \nIgnatov , C. P. Grams , J. Hemberger , I. Nowik , P. S. Halasyamani , T. T. Tran, S. Mukherjee , T. S. \nDasgupta, and M. Greenblatt, Angew. Chem. Int. Ed. 52, 8406 ( 2013 ). \n[17] M.-R. Li, P. W. Stephens , M. Retuerto , T. Sarkar , C. P. Grams , J. Hemberger , M. C. Croft , \nD. Walker , and M. Greenblatt , J. Am. Chem. Soc. 136, 8508 (2014). \n[18] M. Ležaić and N. A. Spaldin, Phys. Rev. B 83, 024410 (2011). \n[19] H. Wang, S. Zhu, X. Ou, and H. Wu, Phys. Rev. B 90, 054406 (2014). \n[20] J. Wang, N. Zu, X. Hao, Y . Xu, Z. Li, Z. Wu, and F. Gao, Phys. Status Solidi RRL 8, No. 9, \n776 (2014). \n[21] H. J. Zhao , W. Ren, Y. Yang , J. Iniguez , X. M. Chen , and L. Bellaiche, Nat. Commun . 5, \n4021 (2014). \n[22] X. Z. Lu, X. G. Gong, and H. J. Xiang , Comput. Mater. Sci. 91, 310 (2014). \n[23] Y. Krockenberger , K. Mogare , M. Reehuis , M. Tovar, M. Jansen , G. Vaitheeswaran , V. \nKanchana , F. Bultmark , A. Delin , F. Wilhelm , A. Rogalev , A. Winkler, and L. Alff, Phys . Rev. \nB 75, 020404 (2007). \n[24] H. L. Feng , M. Arai, Y. Matsushita , Y . Tsujimoto , Y . Guo, C. I. Sathish , X. Wang , Y .-H. \nYuan, M. Tanaka, and K. Yamaura, J. Am. Chem. Soc . 136, 3326 (2014). \n[25] Y. Shimakawa , M. Azuma , and N. Ichikawa, Materials 153, 4 (2011). \n[26] H. J. Xiang, Phys. Rev. B 90, 094108 (2014 ). \n[27] See Supplemental Material for details. \n[28] L. Bellaiche, and D. Vanderbilt, Phys . Rev. Lett. 81, 1318 (1998). \n[29] H. J. Xiang and M. -H. Whangbo, Phys. Rev. B 75, 052407 (2007). \n13 \n [30] H. J. Xiang , C. Lee, H.-J. Koo, X. G. Gong, and M.-H. Whangbo , Dalton Trans. 42, 823 \n(2013); H. J. Xiang , E. J. Kan, S.-H. Wei, M.-H. Whangbo, and X. G. Gong, Phys. Rev. B 84, \n224429 (2011). \n[31] J. B. Goodenough, Phys. Rev. 100, 564 (1955) ; J. Kanamori, J. Phys. Chem. Solids 10, 87 \n(1959). \n[32] The law proposed by Bellaiche et al. [L. Bellaiche, Z . Gui, and I. A. Kornev, J. Phys.: \nCondens. Matter 24, 312201 (2012) ] states that the weak component of the magnetization has a \ndirection that is given by the cross -product of the oxygen octahedral tilting vector (that is lying \nalong the c -direction in the hexagonal setting of the R3 phase) with the G -type antiferromagnetic \nvector (that is chosen here to be along the x -axis). Note that this law also implies that choosing a \nG-type antiferromagnetic v ector being along the same direction as the ox ygen octahedral tilting \nvector (that is, the c -axis) should not produce any weak component of the magnetization \nperpendicularly to the G -type antiferromagnetic vector. This is precisely what we further fou nd \nin the simulations when the Fe and Os moments are chosen to lie (in antiparallel fashion) along \nthe c -axis. \n[33] A. Fert and P. M. Levy, Phys. Rev. Lett. 44, 1538 (1980). \n[34] M. Heide, G. Bihlmayer, S. Blügel, Physica B 404, 2678 (2009) . \n[35] Y . A. Izyumov, Sov . Phys. Usp . 27, 845 (1984) . \n[36] P. S. Wang and H. J. Xiang, Phys. Rev. X 4, 011035 (2014). \n[37] R. E. Cohen, Nature (London) 358, 136 (1992); C. Ederer and N. A. Spaldin, Phys. Rev. B \n74, 024102 (2006). \n[38] J. B. Neaton and K. M. Rabe , Appl. Phys. Lett . 82, 1586 (2003). \n14 \n [39] T. Zhao , A. Scholl , F. Zavaliche , K. Lee, M. Barry , A. Doran , M. P. Cruz , Y . H. Chu, C. \nEderer, N. A. Spaldin, R. R. Das, D. M. Kim, S. H. Baek , C. B. Eom , and R. Ramesh , Nature \nMater. 5, 823 (2006). \n[40] C. Ederer and N . A. Spaldin , Phys. Rev. B 71, 060401 (R) ( 2005 ). \n \nFIG. 1(color online) . (a) The pseudo -cubic structure of double perovskite Zn2FeOsO 6. (b) The \npolar rhombohedral structure of R3 Zn 2FeOsO 6. (c ) The Fe -Os superexchange interactions \nmediated by the corner -sharing O ions (J 1 and J 2) and t he AFM super -superexchange interaction \n(J3 and J 4) between the Os ions . (d) The ferrimagnetic s tructure with the Fe and Os spin moments \naligned in the ab -plane. \n \n15 \n \nFIG. 2(color online). (a) The total energies as a function of the angle α [see the insert ] from the \ndirect DFT+U+SOC calculations (small circles ). The total energy has a minima at αmin≈174° . \nThe total energy curve can be described rather well by the formula \nzz\nspin 1 2 01 04E 3(J J )cos 3(D D )sin \n (blue line) . The insert shows t he DM interaction \nvectors D0i (i =1, 6) between a Fe ion and six Os ions and t he definition of the angl e α between \nthe Fe spin and Os spin in the ab -plane. (b) Relative energies between different spin \nconfigurations. The ferrimagnetic state with the moments along the c -axis is higher in energy by \n0.55 meV/f.u. than the ferrimagnetic state with the Fe and Os moments aligned oppositely in the \nab-plane ( α = 180° ), and t he canting of the spins ( α = 174° ) further lowers the total energy by \n0.97 meV /f.u.. \n \n16 \n \nFIG. 3(color online). The specific heat and t he total in -plane spin moment (M ab) as a function of \ntemperature from the PTMC simulations. The specific heat curve indicates that the ferrimagnetic \nCurie temperature (Tc) is 39 4 K. The in -plane total spin moment increases rapidly near T c. \n \n \n \n \n \n \n17 \n \nFIG. 4(color online). (a) The total energy as a function of the electric polarization for Zn2FeOsO 6. \nIt displays the double well potential with an energy barrier of 0.09 eV/f.u.. (b) Eight possible \norientations of the FE polarization vector ( P) in the sample of double -perovskite Zn2FeOsO 6. (c) \nIllustration of the ME coupling in Zn 2FeOsO 6. A 71° or 109° switch of the FE domains by the \nexternal electric field will be associated with the reorientation of the easy -plane of magnetization. \n \n" }, { "title": "1805.10607v1.Prediction_of_new_multiferroic_and_magnetoelectric_material_Fe3Se4.pdf", "content": "Prediction of new multiferroic and magnetoelectric material Fe 3Se4 \nDeobrat Singh1, Sanjeev K. Gupta2,*, Haiying He3 and Yogesh Sonvane1,* \n1Advance Materials Lab, D epartment of Applied Physics, S. V. National Institute of Technology, \nSurat 395007, India. \n2Computational Materials and Nanoscience Group, D epartment of Physics and Electronics, \nSt. Xavier’s College, Ahmedabad 380009, India. \n3Department of Physics and Astronomy, Valparaiso University, Valparaiso, Indiana 46383, USA . \n \nAbstract \nNowdays , multiferroic materials with magnetoelectric coupling have many real-world applications in \nthe fields of novel m emory devices. It is challenging is to create multiferroic materials with strongly \ncoupled ferroelectric and ferrimagnetic orderings at room te mperature . The single crystal of ferric \nselenide ( Fe3Se4) shows type-II multiferroic due to the coexistence of ferroelectric as well as magnetic \nordering at room temperature. We have investigated the lattice instability , electronic structure , \nferroelectric , ferrimagnet ic ordering and transport properties of ferroelectric metal Fe3Se4. The density \nof states shows considerable hybridization of Fe -3d and Se -4p states near the Fermi level confirming its \nmetallic behavior. The magnetic moments of Fe cations follow a type -II ferr imagnetic and ferroelectric \nordering with a calculated total magnetic moment of 4.25\nB per unit cell (Fe 6Se8). The strong \ncovalent bonding nature of Fe -Se leads to its ferroelectric properties. In addition, the symmetry analysis \nsuggests that tilting of Fe sub -lattice with 3d -t2g orbital ordering is due to the Jahn -Teller (JT) \ndistortion. This study provides further insight in the development of spintronics related technology \nusing multiferroic materials. \n \n*Corre sponding author s: sanjeev.gupta@sxca.edu.in (Dr. Sanjeev Gupta) \nyas@phy.svnit.ac.in (Dr. Yogesh Sonvane) \n \n Introduction \nRecent year, m ultiferroic and magnetoelectric materials are becoming more and more indispensable for \nmany forms of current multi-functional technology, such as highly sensitive magnetic field sensor, \nfilter transducers, filters, phase shifters, memory devices and oscillator s1-4. For practical applications, \nwe needs to discover multiferroic materials at room temperature which is strongly coupled ferroelectric \nand ferrimagnetism ordering. In addition to their important polarization properties, ferroelectrics are \nalso pyroel ectric5, where i t develops a voltage across the material upon heating, while in the case of \npiezoelectric a voltage is developed in response to strain across the material. These properties allow \nferroelectric materials to be utilized in many device applications, including non -volatile memory6, \nthermal detectors7, piezoelectric applications8, and energy harvesters9. The coexistence of magnetism \nand ferroelectricity in a material is called multiferroicity, which is of even great er technological and \nfundamental importance. This has added o ther potential applications . For instance , it can be utilized in \ndata storage system s, which count both magnetic and electr onic state s of the compound to store \ninformation, and in the magneto -elect ronic device s using multiferroic thin films10. \n Some transition metal compounds, such as BiMnO 3 and BiFeO 3 with magnetic Mn3+ and Fe3+ ions, \nare ferroelectric11, owing to the intricate interplay between spin, charge, orbital and lattice degree of \nfreedom in th ese material s12. The mechanisms behind ferroelectricity , however, are not fully \nunderstood because of the complicated structures of early ferroelectrics ( i.e. BaTiO 313). The compound \nCaMnO 3 is also an interesting counter example which shows ferroelectricity and magnetic ordering due \nto the d3 configuration of Mn atom s14. It is worthy to be noted that in general the ferroelectricity \nbehavior require s nearly unoccupied subshell orbitals of the transition metal cations , while partial ly \nfilling d orbitals are required to have magnetic moment s. This dilemma largely hinders the coexistence \nof ferroelectricity and magnetism in a material and explains the scarcity of multiferroic materials15. \n Recently, Bishwas et al.,16 studied the synthesized manganese doped iron selenide nanostructures \nin an attempt t o increase the energy product . Another research group Shao -jie Li et al.,17 obtained high \nCurie temperature and coercivity performance of Fe 3-xCrxSe4 nanostructures which can be utilized in \nthe alternative low -cost hard -magnetic materials . Gen Long et al .,18 demonstrated that at low \ntemperatures, Fe 3Se4 nanostructures exhibit giant coercivity. It was proposed that t his unusual ly large \ncoercivity originates from the large magneto -crystalline anisotropy of the monoclinic structure of \nFe3Se4 with ordered Fe vacancies . The ferromagnetic material BaFe 2Se3 is of particular interest because \nit breaks the parity symmetry of the crystal structure and displ ays exchange striction effects19. Indeed, \nthe iron displacements in the crystal structure are prominent, as reve aled by neutron studies20, 21. Dong \net al.22 show ed that the first nearest -neighbor distances between Fe (↑) and Fe (↑) [or Fe (↓) and Fe (↓)] at 0 K become 3.14 Å, much larger than the Fe (↑) and Fe (↓) distance 2.88 Å. On the other hand, this \nexchange striction is not sufficient to induce Ferroelectric (FE) and Polarization (P) since it breaks \nparity but not space -inversion s ymmetry. Most ferroelectric materials are transition metal oxides having \nvacant d subshells. In this respect the two configurations are not so different , but study of the difference \nin filling of the d -orbitals as theoretically expected for ferroelectricity and magnetism makes these two \nordered states mutually exclusive. According to Giovannetti et. al. [24], the ferroelectric and metallic \nstate are found in LiOsO 3 material. Interestingly, however, i t has been suggested that several other \nmeta llic transi tions could be \"ferroelectric\"23 such as the transition in LiOsO 324 due to the appearance of \na polar axis. \nIn this work, we focus on the Fe3Se4 material derived from the monoclinic phase. We have study \nthe lattice instabilities including ferroelectric distortions in monoclinic Fe 3Se4 materials. We have \nperformed first principles calculations within the spin -polarized density functional theory (DFT) \nframework to obtain the electronic band structure, the role of electron correlations in the metallic state , \nmagnetic properties of the Fe 3Se4 material. \nMethodology \nIn th is work, all calculations are based on the density functional theory (DFT)25 as implemented in \nthe Quantum Espresso (QE) package26. The Kohn -Sham equation s were solved using the Perdew -\nBurke -Ernzerhof (PBE) e xchange -correlation functionals27 formulated within the generalized grad ient \napproximation (GGA) scheme28. We have included sixteen valence electrons for Fe (3s2, 3p6\n, 3d6\n, 4s2) \nand six valence electron s for Se (4s2\n, 4p4) in our calculation s. The kinetic energy cut-off is set to 60 Ry \nwhich yielded good convergence in results of energies and ground state structural parameters. For \nstructural optimization we have used t he conjugate gradient algorithm29. The lattice parameters and the \ninternal coordinates of atoms were optimized with in the space group of I12/m1 (monoclinic ) with the \ncriteria for force and pressure below 10−3 a.u. A Monkhorst -Pack k -point mesh30 of 9×9×11 was used \nfor electronic band structure and density of state calculations , while the spontaneous polarization w as \ncalculated using the Berry phase method31 with a k-point mesh of 5×5×7 . All the calculations were \nspin-polarized . Scalar relativistic Troullier -Martins ultra-soft pseudopotential s32 were employed with \nnon-linear core correc tions33. \n \nResult s and Discussion \nStructural and electronic properties \n \nAs a benchmark test for the approach used in this work, we have investigate d the structural , electronic and magnetic properties of the monoclinic Fe3Se4 phase . Fe3Se4 adopts a normal spinel \nstructure as shown in Fig. 1 and the calculated lattice parameters are listed in Table 1. The Fe3+ ions are \nlocated in the octahedral sites of the monoclinic l attice formed by Se2- anions forming magnetic \nordering . The calculated structural parameter s are in very good agreement with previous reported \nvalues (Table 1). \n \nTable 1. Optimized structural lattice constants of monoclinic Fe3Se4. \nReferenc es a (Å) b (Å) c (Å) β (º) \nPresent 6.100 3.520 11.030 91.15 \nExp34,35,16 6.208, 6.167, 6.159 3.525, 3.537, 3.493 11.2832, 11.170, 12.730 92, 92 , - \n \n \nFigure 1. (A) Relaxed structure of monoclinic Fe3Se4 (Fe in red and Se in yellow) ; (B) distribution of \nelectric dipole moments ( black arrows) of Fe atoms due to partial ionic displacement along the a \ndirection). \n \n \nFigure 2. The electronic band structure of monoclinic Fe3Se4 with up spin (left side) and down spin \n(right side) shown separately. \n \nThe electronic band structure of monoclinic Fe 3Se4 is plotted in Figure 2 . The Fermi level EF is set \nto zero. It is clearly shown that there are bands crossing at EF, thereby the band gap is zero and Fe 3Se4 \nis metallic in nature. This is further confirmed from the density of states ( DOS ) plots (see Figure S1 in \nESI), which demonstrate asymmetric -spin and -spin DOS. A careful look of the projected density \nof states (PDOS) reveals ferrimagnetic spin order ing between the iron and selenium atom s. The \nselenium atoms carried a small magnetic moment (0.14 \nB unit cell ) due to charge transfer and \ngeometric distortion . Therefore , monoclinic Fe3Se4 shows metallic behavior in the ferrimagnetic (FM) \nstate. \nIn order to check whether the metallic behavior is an artifact of neglecting the spin -orbit coupling \n(SOC) , we have further calculated electronic band structure of Fe 3Se4 with the inclusion of SOC \ninteraction (shown in Figure S2 in ESI) . The spin -orbit energy splitting is large r at points of high -\nsymmetry . For instance, t he spin -orbit splitting at gamma point is about 14 meV and 30 m eV in valence \nband maximum (VBM) and conduction band minimum (CBM) , respectively. Since the monoclinic \nFe3Se4 lattice has inversion symmetry, due to the Kramers degeneracy36, each energy band line is at \nleast doubl y degenerate for both spin state s. Our results show that the inclusion of the SOC interaction \ninfluences only three p -energy bands of Se atom s and five d -energy bands of Fe atom s and the metallic \nbehavior remains unchanged . \n \n \nFigure 3. PDOS of Fe ions in relaxed Fe 3Se4. The numbering of Fe is the same as in Figure 1(B). Fe-1. \nFe-2 and Fe -5 forms one layer , while Fe-3, Fe -4 and Fe -6 forms another . \n \n The origin of the metallic behavior of the monoclinic Fe 3Se4 can also be understood from the \nPDOS plots (Figure S1 in ESI). It is clear that the spin -up and spin -down states determining the FM \nmetallicity mainly originate from the edge Fe -3d and edge Se -4p orbitals. Fe atoms contribute more to \nDOS than the Se atoms at Fermi level and the majority of the density of states near the Fermi level for \nFe3Se4 is attributed to the Fe -3d states. The Se -3p bands overlap wit h the Fe -3d bands in the -7 eV to 3 \neV energy range, representing a hybridization of the Se -3p and Fe -3d states to form the covalent \nbonding while Se -3p and Se -3s orbitals also have small contribution to the magnetic property of Fe 3Se4. \nThe difference of t he spin -up band and spin -down band of Fe -3d orbitals show that they carry very \nlarge magnetic moment in Fe 3Se4. Further, the spin polarization is negative near Fermi because of the \nelectronic DOS of spin -down electrons is larger than that of spin -up electr ons. \n Furthermore, we have identified the orbital ordering (OO) state in Fe 3Se4 as shown in projected \nDOS in Figure 3 with five electrons in 3d states of Fe ions in the 3D -coordinates ( xyz) with the x and y \naxes pointing to the crystal [\n101 ] directions and z axis directed to the crystal c-axis. In the minority \nspin channel, four Fe ions (Fe -1, Fe -2, Fe -3 and Fe -4) show t 2g bands right below the Fermi level ( EF) \ndown to1.6 eV with reduced DOS at the EF. All Fe ions in the unit cell of bulk Fe 3Se4 can be \ncategorized into two groups according to the orbital characters : spin-up/down e g bands fully \nunoccupied and spin -down t 2g bands partially occupied, respectively. Bands of Fe -1, Fe -2, Fe -3 and Fe -\n4 [Figure 3(A, B, C and D)] are of predominate d x2\n-y2and d z2 characters, while bands of Fe -5 and Fe -6 \n[Figure 3(E) and 3(F)] are of mainly d xy, dz2 and d x2\n-y2 orbitals. \nThese result s demonstrate that the configuration of -spin 3dt 2g OO state s in the Fe sub -lattice , \nwhere a -spin electron of each Fe has occup ied a mixed t 2g state. The mixing of orbitals is a \ncombination of two of the t 2g states with the third empty t 2g states37. In addition, it is found that the -\nspin electrons of Fe from 3dt 2g OO state occupying, respectively, the canted d xy and d yz orbitals of Fe -1 \nto Fe -4 and d yz of Fe-5 and Fe -6 sublattice , as they contribute more at the Fermi level . The OO pattern \nis clearly seen in the distribution of charge density on each Fe3+ with the distribution on Fe -1, Fe -2, Fe -\n3 and Fe -4 belonging to one state, while with Fe -5 and Fe -6 belonging to another state. \n \nSymmetry analysis \nAccording to the ionic model37, we have acquired a metallic ground state in consistence . It is also \nshown that the tilting 3dt 2g OO states on the Fe sub -lattice is strongly related to the John -Teller (JT) \ndistortions37. The main contributions in electronic structure were utilized for Fe ions to investig ate the correlation effects in 3d -electrons. The OO states are generally found in 3de g manganite frameworks \nwhere the helpful Jahn -Teller (JT) mutilations are huge because of the solid hybridization between the \neg and O -2p electrons38. The t2g OO states with higher degeneracy and moderately weaker JT distortion \nis additionally found in confined 3d frameworks magnetite39. In the ionic model, the five 3d electrons \nof the Fe3+ ion possess the t 2g triplet degenerate and leave the higher e g doublet degenerate vacant, \nunder the octahedral crystal field (see Figure S4 in ESI) . As per Hund's rule , Fe3+ is in the high spin \nstates with the spin arrangement of (\n3\n2gt\n2\n2gt ), giving rise to a magnetic moment of 4.25\nB unit \ncell: and it indicates metallic ground state with the majority -spin at the Fermi level accountable for \nthe conductivity . Assess ing three 3d electrons, the spin up t 2g states is moderately confined with a \nsuppressed bandwidt h and energy of lower band, while the -spin t2g band is pushed upwards \nmarginally [se e ESI in Figure S 1]. Besides, the -spin t2g band is thus completely occupied and at E F \nthere is no band gap , giving rise to a metallic ground state in consistence with the valence setup (\n3\n2gt\n2\n2gt\n ) of Fe from ion ic model as display ed in Figure S 3 in ESI . \nThe tilting of FeSe 6 octahedral site (Fig. 1 (A)) is to accommodate two sorts of distortion of the Se \nlattice : the two Se ions at the tip of the octahedron deform along the z-axis (Fe-Se top bond length = \n2.44 Å), while two ions of the coplanar Se move upwards and two ions move downwards along the z-\naxis. The JT distortions with Fe-Se bond length (2.46 Å , between Fe -1 and Se ) along the y-axis and \nextended bond length (2.63 Å) along the x-axis (Figure 1) split the triply degenerate t2g orbitals into the \nlower d xz and higher d yz and d xy orbitals . The JT distortions could bring down the Coulomb interaction \nbetween the ions of Se and the Fe d xz states , which is a combination of d xy and d yz states as appeared in \nFigure S 4 in ESI . Moreover , the d xz states with intermediate directions between nearest Se anions would \nadditionally stabilize the lattice distortion. \n It is well known tha t, if we include the spin -orbit coupling (SOC) interaction then the observed \norbital -ordering in SrRuO 3 would be destroyed ; otherwise if we do not include SOC interaction then \nthe polyhedral crystal field makes the cubic harmonics (break ing the inversion symmetry)40 a nature \nbasis set resulting in lower t 2g and higher e g bands37. The Se octahedral site also break parity in each \nsite as Figure 1 shows that Se -5 is above the ladder’s plane, but the next Se -7 is below, and the \ndistances of Se -5 and Se -7 to the iron ladder plane should be the same in magnitude and opposite in \nsign (“antisym metric”). However, the OO introduces a fundamental modification in the symmetry. \nNow the blocks m ade of four Fe (↑) [or four Fe (↓)] are no longer identical to pair of two Fe (↑) and \ntwo Fe (↓). Then, the heights of Se-5 and Se-7 do not require to be antisymmetric anymore; their \ndistances to the xz plane can become different. A similar mechanism works for the edge Se's, e.g., Se -1 and Se -5. As a result, the atomic positions of Se break the space inversion symmetry, creating a local \nFerroelectric (FE) Polarization (P) pointing perpendicular to the iron yz plane (almost along the x- axis). \nTo clarify more of the features of OO states, we have calculated charge density distribution (0.66e/ \nÅ3) and depicted in Figure S5 in ESI which corresponds to the -spin t2g bands below the Fermi level ( -\n1.80-0.0 eV) [see ESI in Figure S1]. It has been observed that electron de nsities corresponding t o -\nspin electrons are totally different for each atom present in the zigzag whereas when we compare both \nthe spins, electron densities are partially different from each other. This suggests that in case of Fe 3Se4, \nspin down electrons present at unsaturated edge Fe -atoms are responsible for conduction as also \nconfirmed through corresponding band structure and DOS. The charge contour of real space \ndistribution of spin dependent electron densities suggests that the possibility of metallic spin \npolar ization in Fe 3Se4 structures is totally attributed to the localization of unpaired electrons at \nunsaturated Fe atoms which are situated at face s of the unit cell in Figure S 5(A) and S5(B) in ESI . \n \n The yellow spheres show that the Se anions attract more electrons while, the Fe cations lose more \n3d electrons in Fe 3Se4, shown by bright magenta lobes pointing along the Fe directions. By contrast, the \ndensity difference is weak but also exists in the Se ladder plane. There is a dark yellow sphere \nintroduce d as a Se atom, with negative value: this suggests that the outmost electrons of Fe are more \nextended ( delocalized), also supporting the covalent scenario for Fe 3Se4. \n \nMagnetic and ferroelectric properties \n \nThe calculated total magnetic moment per unit cell o f monoclinic Fe3Se4 with 14 atoms per unit cell \n(6-Fe atom and 8-Se atom) is 4.25\nB . This value is consistent with the previous DFT calculations \n(4.34\nB /unit cell)35, but significantly higher than the reported experimental value for \nFe3Se4 nanostructures (2.2\nB /unit cell )35. This was attributed to the spin fluctuation and the long -range \nordering as measured by experiments, which normally would be smaller than the calculated value at 0 \nK. \nWe have calculated spontaneous polarization (Ps) along each of the directions x, y and z, for the \nferroelectric phase apply ing the Berry phase approach for bulk Fe3Se4. The direction of Ps in Fe 3Se4 \nsingle crystal lies in the monoclinic ac plane and the magnitude of the segment of spontaneous \npolarization along the a-axis of the monoclinic un it cell. The spontaneous polarization (Ps) in Fe 3Se4 \nhas a value of ~44.60 μC/cm2 and the direction of Ps vector makes an angle 91.15° with the major \nsurface (normal to c-direction). The calculated values of three component s of the spontaneous polarization vectors are P x= 23.14 μC/cm2, Py= 6.20 μC/cm2and P z=37.62 μC/cm2, respectively. The \nmagnitude of spontaneous polarization of Fe 3Se4 is in very good agreement with previously reported \nexperimental value41. Roy et al .,42 provided the spontaneous polarization of ferroelectric bismuth \ntitanate with a value of 42.83 μC/cm2 which is fairly comparable to our calculated value for \npolarization . Another study by Ravindran et al.,43 gave the spontaneous polarization of BiFeO 3 which is \nalso within the reported agreement between theoretical and experimental values. Therefore, it is clear \nthat the calculated values of spontaneous polarization using GGA fall in the range of 43 -68 μC/cm2. \nDielectric and transport properties \n Furthe rmore, we have calculated the frequency dependent optical properties including dielectric \nfunction, absorption coefficient using DFT within the random phase approximation (RPA)44. The \nfrequency dependent dielectric function can be written as \n12 ( ) ( ) ( ) i . Where\n1() and \n2()\nare the real and imaginary parts of the complex dielectric function, respectively. \n2() is \ndetermined by summation over electronic states and \n1() is obtained using the Kramers –Kronig (KK) \nrelationship45. \n The real and imaginary parts of the complex dielectric function \n versus frequency for bulk \nFe3Se4 are presented in Figure (4A and 4B). The real (\n1 ) and imaginary (\n2 ) part follow the general \ntendency of a metal which can be well explained by the classical Drude theory. In Figure 4A and 4B, \nwith increasing in frequency, the\n1 decreases while \n2 initially increases and decreases above the \n1.5x1014 Hz up to 2.0x1014 Hz region, as expected in a metal46. The variation of dielectric constant \n \nand loss tangent is shown in Figure 4C and 4D. With increasing frequency, the \n decreases and \ntan\ninitially increases at certain frequencies and then it is decreases in all the polarization direction (in -\nplane and out -of-plane). The loss tangent measures the loss -rate of power in an oscillatory dissipative \nsystem47. It is clearly seen that the variation \ntan executes in the same trend as \n2 . Since \n2 is lower \nthan \n1 , then the energy loss of the materials is relatively low. This suggests that the material possesses \ngood optical qualities due to lower energy losses and lower scattering of the incident radiation47. Our \ntheoretical results of the dielectric function of bulk Fe 3Se4 are in excellent agreement with previously \nreported experimental results48 except \n2 . But according to other theoretical investigation our results is \nexpected in a metal46. \n \nFigure 4. (A) Real and (B) imaginary part of the complex dielectric constant, (C) magnitude of the \ncomplex dielectric function and (D) loss tangent verses frequency at 1014 Hz at room temperature (300 \nK) of bulk Fe 3Se4. (E) negative value of real part of complex dielectric function is a function of photon \nenergy. \n The real part of complex dielectric function \n1 is shown in Figure 4(E). In Fig ure 4(E) , we plot \nthe real parts of dielectric functions in all the polarisation direction for the materials. Character of this \nmaterial exhibit metallic behavior, i.e. Drude peaks at low energy due to intraband contribution and the \nreal part of ε(ω) crossing from negative value to positive value with increasin g frequency and it’s \nshows o scillatory behavior upto 15 eV . The strong anisotropy in the real part of complex dielectric \nfunction is also quite obvious. Remarkably, because of this anisotropy, the real part of complex \ndielectric function, \n1 in x, y and z -direction , changes sign at a frequency which is different from \neach other , leading to an extended frequency window in which the each components have different \nsigns shown in shaded region with gray color in Figure 4(E) . Such type of sign difference is the \ncharacteristic feature in real part of dielectric function is known as indefinite media49. Generally, \nindefinite media are mostly in artificially assembled structures which require complicated fabrication \nprocess and usua lly have high dissipation. This theoretical results suggest that crystalline solid Fe 3Se4 \nin the bulk form would just be indefinite materials for a frequency range spanning the near infrared. \n \nFigure 5. The variation of (A) thermal conductivity\n , (B) heat capacity \npC as a function of \ntemperature. \n The thermal conductivity ( ) of a material originates from two fundamental sources: (i) electrons \nand hole s transporting heat (κ e) and (ii) phonons trave lling through the lattice (κ l). Most of the \nelectronic term (κ e) is straight forwardly associated with the electrical conductivity which can be well \nexplained by Wiedemann –Franz hypothesis50. From Figure 5(A), the thermal conductivity of Fe3Se4 is \nalmost linearly increased from 50 K to 400 K , and at room temperature the thermal conductivity is 4.56 \nW/ (m K). The thermal conductivity is dominated by the lattice contribution as it is about 1.0 W/m K at \nT=50 K and it increases to 5.57 W/mK at T=400 K which is higher than the values observed in most \nthermoelectric materials50, 51. Since the heat flow in a material is directly proportional to its thermal \nconductivity , the heat flow in bulk Fe 3Se4 will be higher as well . \n Specific heat estimation is one of the most reliable techniques for exploring temperature reliance of \nmaterials. The heat capacity \npC of bulk Fe 3Se4 increas es with temperature as shown in Figure 5(B ), \ndemonstrat ing a similar behavior as the thermal conductivity . In addition, the variation of \npC with \ntemperature follow s the well-known Debye form and shows excellent fit at lower temperatures. Our \ntheoretical results are in good agr eement with experimental results48. Furthermore, the estimated \nelectric conductivity as a function of temperature of Fe 3Se4 reveals nearly a linear behavior with a \npositive slope as plotted in Figure 6. This could be ascribed to the way that the electrical conductivity \nwas evaluated based on DFT by assuming a constant scattering time \n14( 10 ) s . In general, in ca se of \nmetallic system s, the electrical conductivity decreases with respect to temperature owing to the \noccurance of more collisions among electrons and between electrons and phonons which shorten the \nmean free path of charge carriers. I n our case , similar behavior are found in its metallic system , the \nelectrical conductivity of Fe 3Se4 continuously decreases with increasing tem perature. This is likely due \nto the enhanced delocalization of hybrid orbitals. \n \nFigure 6. The variation of electronic conductivity\n as a function of temperature. \n \n \n \n \n \n \n \n \n \n \n \nFigure 7. The absorption coefficient \n in the unit of 105/cm of bulk Fe 3Se4 for all the polarization \ndirection in the electric field. (A) The absorption coefficient up to 15 eV and (B) absorption coefficient \nup to 40 eV beyond 28 eV the absorption will be almost zero . \nThe absorption coefficient \n in all the polarization directions of the electric field (depicted in \nFigure 7) show two main peaks in all polarizations. The first main peak oc curs at energy around 5 eV \nthat is related to π electron plasmon ic peak s. And other peak occurs around 15 eV that is associated to \nπ+σ electron plasmon ic peak s. Moreov er, at the energy range of 10 -20 eV, the value of absorption for \nall cases is very high. \nConclusion s \nIn conclusion s, we investigated the novel multiferroic material shows ferroelectric, ferrima gnetism \nand electronic structure in monoclinic Fe3Se4 material. The nature of orbital ordering , and its close \nconnection to the JT distortion is unwound, which is responsible for ferroelectric like instability . Fe3Se4 \ntakes a ferromagnetic ordering with a total magnetic moment of 4.25\nB per unit cell (6 Fe atom and 8 \nSe atom) . The spontaneous polarization (44.60 μC/cm2) provides evidence of its ferroelectric behavior. \nOur study suggests that the monoclinic phase of Fe 3Se4 material possesses the unusual dual behavior of \nmetallicity and ferroelectricity . The stabilization of the ferroelectric structure in Fe 3Se4 coexisting with \nmetallic conductance is the consequence of a decoupling between the metallic electrons in the t 2g \nelectrons from the soft phonons which break the inversion symmetry . The development of multiferroic \nmaterial are usefull applications in spintronics -related technologies for ultrahigh -density memory and \nquantum -computer devices is underway. \n \nAcknowledgem ents \nHelpful discussion with Drs. Pankaj Poddar and Mousumi Sen is acknowledged . S. K. G. \nacknowledges the use of high performance computing clusters at IUAC, New Delhi and YUVA, \nPARAM II, Pune to obtain the partial results presented in this paper. S . K. G and Y. A. S also thank the \nScience and Engineering Research Board (SERB), India for the financial support (grant nu mbers.: \nYSS/2015/001269 and EEQ/2016/000217, respectively ). D. S. would like to thank University Grant \nCommission (UGC), New Delhi, India for the financial support. \n \nReferences \n \n1. C. W. Nan, M. I. Bichurin, S. Dong , D. Viehland and G. Srinivasan, Multiferroicmagnetoelectric \ncomposites: historical perspective, status, and future directions, J Appl. Phys. , 103(3) , 031101 \n(2008) . \n2. J. Zhai , S. Dong, Z. Xing, J. Li and D. Viehland, Geomagnetic sensor based on giant \nmagnetoelectric effect, Appl. Phys. Lett. , 91, 123513 (2007) . \n3. M. Bichurin, R. Petrov and Y. V. Kiliba, Magnetoelectric microwave phase shifters, Ferroelectrics , \n204(1) , 311 -319 (1997) . \n4. Y. K. Fetisov and G. Srinivasan, Electric field tuning characteristics of a ferrite -piezoelectric \nmicrowave resonator, Appl. Phys. Lett. , 88(14) , 143503 (2006) . \n5. R. Whatmore, Pyroelectric devices and materials, Rep. Prog. Phys. , 49(12) , 1335 (1986) . \n6. Y. Wang, J. Hu, Y. Lin, C. W. Nan, Multiferroic magnetoelectric composite nanostructures, NPG \nAsia Materials , 2(2), 61 -68 (2010) . \n7. D. McCammon, R. Almy, E. E. A. Apodaca, W. B. Tiest, W. Cui, S. Deiker, M. Galeazzi, M. Juda, \nA. Lesser and T. Mihara, A hig h spectral resolution observation of the soft X -ray diffuse \nbackground with thermal detectors, Astrophys. J. , 576(1) , 188 (2002) . \n8. K. Shimamura, H. Takeda, T. Kohno and T. Fukuda, Growth and characterization of lanthanum \ngallium silicate La 3Ga5SiO 14 single crystals for piezoelectric applications J. Cryst. Growth, 163(4) , \n388-392 (1996) . \n9. A. Erturk and D. J. Inman, A distributed parameter electromechanical model for cantilevered \npiezoelectric energy harvesters. J. Vib. Acoust. , 130(4) , 041002 (2008) . \n10. M. Vopsaroiu, J. Blackburn and M. G. Cain, A new magnetic recording read head technology based \non the magneto -electric effect, J. Phys. D Appl. Phys. , 40(17) , 5027 (2007) . \n11. S.-W. Cheong and M. Mostovoy, Multiferroics: a magnetic twist for ferroelectricity, Nat. Mater. \n6(1), 13-20 (2007). \n12. S. Liang, A. Moreo and E. Dagotto, Nematic state of pnictides stabilized by interplay between spin, \norbital, and lattice degrees of freedom, Phys. Rev. Lett. , 111(4) , 047004 (2013) . \n13. Z.-X. Chen, Y. Chen, and Y. -S. Jiang, DFT study on ferroelectricity of BaTiO 3, J. Phys. Chem. B , \n105(24) , 5766 -5771 (2001) . \n14. C. Ederer, T. Harris, and R. Kováčik, Mechanism of ferroelectric instabilities in non -d0 \nperovskites: LaCrO 3 versus CaMnO 3, Phys. Rev. B , 83, 054110 (2011) . \n15. B. Wul and I. Goldman, Dielectric constants of titanates of metals of the second group, Dokl. Akad. \nNauk SSSR , 46, 139 –142 (1945) . \n16. M. Sen Bishwas, R. Das, and P. Poddar, Large increase in the energy product of Fe 3Se4 by Fe -site \ndoping, J. Phys. Chem. C , 118(8) , 4016 -4022 (2014) . 17. S.-j. Li, D. Li, W. Liu, and Z. Zhang, High Curie temperature and coercivity performance of Fe 3− x \nCrxSe4 nanostructures, Nanoscale , 7(12) , 5395 -5402 (2015) . \n18. G. Long, H. Zhang, D. Li, R. Sabirianov, Z. Zhang, and H. Zeng, Magnetic anisotr opy and \ncoercivity of Fe 3Se4 nanostructures, Appl. Phys. Lett. , 99(20) , 202103 (2011) . \n19. S. Dong, J. -M. Liu, and E. Dagotto, BaFe 2Se3: A High TC Magnetic Multiferroic with Large \nFerrielectric Polarization, Phys. Rev. Lett. , 113(18) , 187204 (2014). \n20. J. Caron, J. Neilson, D. Miller, A. Llobet, and T. McQueen, Iron displacements and magnetoelastic \ncoupling in the antiferromagnetic spin -ladder compound BaFe 2Se3, Phys. Rev. B , 84, 180409 \n(2011) . \n21. B. Saparov , S. Calder, B. Sipos, H. Cao, S. Chi, D. J. Singh, A. D. Christianson, M. D. Lumsden, \nand A. S. Sefat, Spin glass and semiconducting behavior in one -dimensional BaFe 2−δSe3 (δ≈ 0.2) \ncrystals, Phys. Rev. B , 84, 245132 (2011) . \n22. P. S. Wang, W. Ren, L. Bellaich e, and H. J. Xiang, Predicting a ferrimagnetic phase of Zn2FeOsO \n6 with strong magnetoelectric coupling. Phys . Rev. Lett., 114(14) , 147204 (2015) . \n23. P. W. Anderson and E. Blount, Symmetry Considerations on Martensitic Transformations: \n“Ferroelectric” Metals?, Phys. Rev. Lett. , 14 (13) , 532 –532 (1965) . \n24. Y. Shi, Y. Guo, X. Wang, A. J. Princep, D. Khalyavin, P. Manuel, Y. Michiue, A. Sato, K. Tsuda, \nand S. Yu, Ferroelectric -like Struct ural Transition in a Metal, Nat. Mater. , 12 (11) , 1024 –1027 \n(2013) . \n25. G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total -energy calculations \nusing a plane -wave basis set, Phys. Rev. B , 54(16) , 11169 (1996) . \n26. P. Giannozzi, S. Baroni , N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, \nM. Cococcioni, and I. Dabo, QUANTUM ESPRESSO: a modular and open -source software project \nfor quantum simulations of materials, J. Phys. Condens. Matter. , 21, 395502 (2009) . \n27. J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized Gradient Approximation Made Simple, \nPhys. Rev. Lett, 77, 3865 -3868 (1996) . \n28. (a) J. P. Perdew and Y. Wang, Accurate and simple analytic representation of the electron -gas \ncorrelation energy, Phys. Rev. B , 45(23) , 13244 (1992) . (b) J. P. Perdew, K. Burke, and Y. Wang, \nGeneralized gradient approximation for the exchange -correlation hole of a many -electron system, \nPhys. Rev. B , 54(23) ,16533 (1996) . \n29. W. H. Press, B. Flannery, S. Teukolsky, and W. Vetterling, Num erical Recipes. New York: \nCambridge University Press , 1986. \n30. H. J. Monkhorst and J. D. Pack, Special points for Brillouin -zone integrations, Phys. Rev. B , 13, \n5188 -5192 (1976) . \n31. D. Vanderbilt and R. King -Smith, Electric polarization as a bulk quantity and it s relation to surface \ncharge, Phys. Rev. B , 48, 4442 –55 (1993) . \n32. D. Hamann, M. Schlüter, and C. Chiang, Norm -conserving pseudopotentials, Phys. Rev. Lett. , 43, \n1494 -1497 (1979) . \n33. S. G. Louie, S. Froyen, and M. L. Cohen, Nonlinear ionic pseudopotentials in spin -density -\nfunctional calculations, Phys. Rev. B , 26, 1738 -1742 (1982) . \n34. C.-R. Lin, Y. -J. Siao, S. -Z. Lu, and C. Gau, Magnetic properties of iron selenide nanocrystals \nsynthesized by the thermal decomposition, IEEE Tran s. Magn. , 45, 4275−4278 (2009) . \n35. G. Long, H. Zhang, D. Li, R. Sabirianov, Z. Zhang, and H. Zeng, Magnetic anisotropy and \ncoercivity of Fe 3Se4 nanostructures, Appl. Phys. Lett. , 99, 202103 (2011) . \n36. M. Tinkham, Group Theory and Quantum Mechanics. McGraw -Hill ( New York) , 1971. \n37. H.-T. Jeng, S. -H. Lin, and C. -S. Hsue, Orbital ordering and Jahn -Teller distortion in Perovskite \nruthenate SrRuO 3, Phys. Rev. Lett. 97(6) , 067002 (2006) . \n38. Y. Tokura and N. Nagaosa, Orbital physics in transition -metal oxides, Science , 288, 462 (2000) . 39. H.-T. Jeng, G. Guo, and D. Huang, Charge -orbital ordering and Verwey transition in magnetite, \nPhys. Rev. Lett., 93, 156403 (2004) . \n40. M. Heide, G. Bihlmayer, P. Mavropoulos, A. Bringer, and S. Blügel, Spin Orbit Driven Physics at \nSurfaces, Newsletter of the Psi -K Network, 78 (2006) . \n41. S. Cummins and L. Cross, Electrical and Optical Properties of Ferroelectric Bi 4Ti3O12 Single \nCrystals, J. Appl. Phys, 39(5) , 2268 -2274 (1968) . \n42. A. Roy, R. Prasad, S. Auluck, and A. Garg, First-principles calculati ons of Born effective charges \nand spontaneous polarization of ferroelectric bismuth titanate, J. Phys. Cond. Matt., 22(16), \n165902 -165926 (2010) . \n43. P. Ravindran, R. Vidya, A. Kjekshus, H. Fjellvåg, and O. Eriksson, Theoretical investigation of \nmagnetoelectri c behavior in BiFeO 3, Phys. Rev. B, 74(22) , 224412 -224418 (2006) . \n44. M. Gajdoš, K. Hummer, G. Kresse, J. Furthmüller, and F. Bechstedt, Linear optical properties in \nthe projector -augmented wave methodology, Phys. Rev. B: Condens. Matter. Mater. Phys. , 73, \n045112 (2006) . \n45. M. Fox, Optical Properties of Solids. Oxford University Press(New York) , 2001, 3. \n46. B. Mortazavi, M. Shahrokhi, M. Makaremi, and T. Rabczuk, Anisotropic mechanical and optical \nresponse and negative Poisson's ratio in Mo 2C nanomembranes revealed by first -principles \nsimulations, Nanotech. , 28(11) , 115705 (2017) . \n47. M. M. Rahman, H. A. Miran, Z. -T. Jiang, M. Altarawneh, L. S. Chuah, H. -L. Lee, A. Amri, N. \nMondinos, and B. Z. Dlugogorski, Investigation of the post -annealing electromagnetic response of \nCu–Co oxide coatings via optical measurement and computational modelling, RSC Adv. , 7(27) , \n16826 -16835 (2017) . \n48. M. S. Bishwas and P. Poddar, Discovery of room temperature multiferroicity and magneto -electric \ncoupling in Fe 3Se4 nanorods, arXiv preprint arXiv: 1612.06512 , (2016 ). \n49. D. Smith and D. Schurig, Electromagnetic Wave Propagation in Media with Indefinite Permittivity \nand Permeability Tensors, Phys. Rev. Lett. , 90, 077405 (2003) . \n50. G. J. Snyder and E. S. Toberer, Complex thermoelectric materials, Nat. Mater. , 7(2), 105 -114 \n(2008) . \n51. J. R. Sootsman, D. Y. Chung, and M. G. Kanatzidis, New and old concepts in thermoelectric \nmaterials, Angew. Chem. Int. Ed. , 48(46) , 8616 -8639 (2009) . " }, { "title": "2004.11406v2.Temperature_dependent_Magnetic_Transitions_in_CoCrPt_Ru_CoCrPt_Synthetic_Ferrimagnets.pdf", "content": " 1 Title: Temperature -dependent Magnetic Transitions in CoCrPt -Ru-CoCrPt Synthetic Ferrimagnet s \nAuthors: Bradlee Beauchamp and Ernesto E. Marinero \nSchool of Materials Engineering, Neil Armstrong Hall of Engineering, Purdue University, 701 W Stadium Ave, West \nLafayette, IN 47907 \nCorresponding Author: eemarinero@purdue.edu \nDOI: https://doi.org/10.1016/j.jmmm.2020.167586 \n© 2020. This manuscript version is made available under the CC -BY-NC-ND 4.0 license \nhttp://creativecommons.org/licenses/by -nc-nd/4.0/ \nAbstract: \n The magnetic orientations and switching fields of a CoCrPt -Ru-CoCrPt synthetic ferrimagnet with \nperpendicular magnetic anisotropy have been studied in the temperature range from 2 K to 300 K. It was \nfound that two sets of magnetic transitions occur in the CoCrPt -Ru-CoCrPt ferrimagnet across this \ntemperature range . The first set exhibits three magnetic transitions in the 50 K – 370 K range, whereas the \nsecond involves only two transitions in the 2 K and 50 K range. The observed magnetic hysteresis curves of \nthe synthetic ferrimagnet are explained using the energy diagram technique framework pioneered by Koplak \net al. [1] which accurately describes the competition between interlayer exchange coupling energy, Zeeman \nenergy, and anisotropy energy in the system. In this work we expand the framework to include synthetic \nferrimagnets ( SFMs ) comprising higher perpendicular magnetic anisotropy materials and large (4X) interlayer \nexchange coupling energies which are promising for the development of ultrafast (ps) magnetic switching free \nlayers in MTJ structures . Furthermore, we apply the analysis to predict SFM magnetic hysteresis curve s in a \ntemperature regime that includes temperature extrema that a synthet ic ferrimagnet would be expected to \nreliably operate at, were it to be utilized as a free layer in a memory or sensor spintronic device. \nKeywords: \nSynthetic ferrimagnet, Magnetic films , CoCrPt , Interlayer e xchange coupling, Perpendicular magnetic \nrecording, Magnetic switching , spintronics, magnetic tunnel junctions \nIntroduction: \nSynthetic ferrimagnet ( SFM ) trilayers consist of two antiparallel ferromagnetic (FM) films separated by \na thin non-ferromagnetic metallic interlayer. For the case of identical FM layers, if the films are dissimilar in \nthickness, the SFM structure will exhibit a net magnetic moment (uncompensated ferrimagnet) . The interlayer \nexchange coupling energy (IECE) of the SFM varies with the interlayer thickness in an oscillatory fashion [2] \nand it has been attributed to various physical processes that include dipolar magnetostatic interaction s and 2 Ruderman -Kittel -Kasuya -Yosida (RKKY) coup ling. First observed by Gr ünberg et al. [3], films exhibiting \nantiferromagnetic coupling were utilized shortly thereafter in magnetic sen sor devices based on the giant \nmagnetoresistance (GMR) observed in Fe/Cr antiferromagnet s tructures [4][5]. More recently, SFMs and \nantiferromagnets have been utilized in magnetic tunnel junctions (MTJs) to provide exchange bias to the \nreference layer (typically a single, magnetically hard ferromagnet) such that its magnetization is “fixed” [6,7] . \nSFM trilayers or antiferromagnets have been considered to replace the recording or reference layer of the MTJ \n[8,9] . When utilized in an MTJ device the strength of coupling can determine if the SF M is acting as a reference \nor free layer . The coupling strength is derived from measuring the magnetic field required for ov ercoming the \nIECE which renders the magnetization orientation of the individual layers to be parallel. It is noted that free \nlayer SFM structures incorporated into MTJ devices have demonstrated lower critical switching current s than \nsingle FM free layers with out negatively affecting thermal stability [10]. Additionally , we have proposed that \nSFM free layers can exhibit ultrafast switching speeds down to the picosecond time regime [11]. \nMost MTJ devices utilize CoFeB as the FM electrode due to high tunneling magnetoresistance \nmeasured when used with MgO tunneling barrier s [12]. However, the maximum thickness of CoFeB exhibiting \nperpendicular magnetic anisotropy (PMA ) is limited to around 1 .5 nm [13]. The magneti zation of CoFeB is also \nrelatively high, which increases the charge current needed for spin -transfer torque switching. CoCrPt is a \nmaterial of interest for MTJ applications due to its low magnetization and its large anis otropy [14], resulting in \nlower switching current s, improve d thermal stability , and the use of thicker FM layers with concomitant \nprocess control improvements. In addition, the SFM configuration circumvents the materials -restrictive low \nmagnetic damping requirement for selection of the FM thin film for MTJ devices [11]. \n It is essential to tailor the IECE and switching properties of the S FM structure for use in memory \ndevices . However, the magnetic properties of the SFM are temperature dependent , and memory devices \ncould be expected to operate under extreme conditions within the range of 2 00 K to 370 K. In this paper the \nIECE of CoCrPt -Ru-CoCrPt trilayer structures has been investigated from 2 K to 300 K. It has been observed by \nKoplak et al . [1] that with decreasing temperature , the hysteresis loops of S FMs vary dramatically , and these \nauthors developed a formalism to describe the observed changes in the hysteresis loop as a function of \ntemperature in a CoFeB -Ta-CoFeB synthetic antiferromagnet . They employ an energy balance approach that \nincludes the Zeeman energy, the IECE, and energy barriers for switching arising from the effective magnetic \nanisotropy energy. It was found that t he two main parameters controlling the switching behavior with \ndecreasing temperature is the ratio of the magnetic moments of the two constituent ferromagnetic layers as \nwell as the energy barrier for switching of each film, which is temperature dependent . In this paper th e energy \ndiagram technique introduced by Koplak et al. is used to describe the magnetic transitions measured in a 3 CoCrPt -Ru-CoCrPt SFM as a function of temperature. This is compared with their results on the CoFeB -Ta-\nCoFeB S FM structure. Predictions are also made for the magnetic transitions of the CoCrPt -Ru-CoCrPt SFM at \n200 K to 370 K to exemplify the practical use of the energy diagram technique for assessing the robustness of \na potential sensor or memory device employing a SFM read layer. \n \nMaterials and Methods : \nAll films were deposited without substrate heating or bias in a magnetron sputter system with a base \npressure < 10-7 Torr. The films were grown on oxidized silicon (100) substrates, with dimensions ~ 5mmx5mm. \nThese small coupons were obtained by manual dicing of larger (2” diameter) Si wafers. Thus, the actual \ndimensions of each sample exhibited small variations. Area measurem ents for each coupon were performed \nwith optical microscopy with a reference scale. The error in surface area determination is ~8%. The thin film \nstructure consisted of the following: Ta(5 nm)/Ru(10 nm)/CoCrPt( 1.7 nm)/Ru( X nm)/CoCrPt( 1.3 nm)/Ru(5 nm). \nThe CoCrPt sputtering target has a nominal composition of Co 70Cr18Pt12. The Ta/Ru seed layer was used to \npromote CoCrPt growth with its basal (002) plane parallel to the thin film plane (c -axis out of plane) . Magnetic \nhysteresis loops were collected using a Quantum Design MPMS -3 superconducting quantum interference \ndevice (SQUID) magnetometer with 10-8 emu sensitivity. All samples were mounted in the magnetometer \nusing a plastic straw attachment. Measurements at 370 K could not be performed using this method d ue to \nerrors on account of warping of the straw mount at this temperature. All magnetic hysteresis loops presented \nin this work were performed with the substrate aligned perpendicular to the direction of the applied magnetic \nfield. \nResults and Discussion: \n \nFigure 1 . Left: Schematic representation of the film stack cross -sectio n: red arrows indicat e the direction of \n \n \n \n \n \n \n 4 magnetization of the constituent FM layers at remanence . Right: Hysteresis curves of the S FM structures with \ntwo different Ru interlayer thicknesses measured at 300 K. \n As shown in Fig. 1, the IECE of the CoCrPt -Ru-CoCrPt S FM can be tailored by varying the Ru interlayer \nthickness. The IECE per unit area of a S FM with dissimilar FM layers was esti mated by Koplak et al. using the \nexpression: JEX = -HBm2/S. Here H B is the bias field which indicates the center of the outer loop . In Fig. 2a-b, H B \nis labeled and measured by finding the center of the outer loops at which point the SFM becomes saturated. \nAt 2 K (Fig. 2c), there is no outer loop unlike in Fig. 2a -b. In this case a minor loop must be taken to locate the \nHB field, shown by the red curve in Fig. 2c . Peng et al. [15] performed similar minor loop measur ement to \ndetermine the exchange coupling field in CoCrPt (18 nm)/Co(1 nm)/Ru(0.9 nm)/CoPtCr(8 nm) A FM structures \nas the magnetic measurements of these samples exhibited no outer loops present in the major loop of the \nhysteresis curve. A comparable measureme nt was later performed by Bandiera et al. [9], where a minor loop \nwas measured to deduce the exchange coupling field in a SFM structure comprising Co/Pt multilayers , Co \ninterlayers, Ru and CoFeB. Both studies demonstrate that the exchange coupling field ca n be obtained by the \nreversal of the softer magnetic layer in the SFM. Peng et al. measured this minor loop in the first quadrant of \nthe hysteresis loop, while Bandiera et al. measure d the minor loop in the third quadrant . \nIn the J EX equation, m2 is the magnetic moment from the thinner magnetic layer (m1 being the moment from \nthe thicker layer) , and S is the surface area of the film. The measured surface areas, S, are determined to be \n0.21 ± 0.01 cm2 and 0.09 ± 0.01 cm2 for the hyst eresis loops shown in Fig. 1 for the SFMs with 0.5 nm and 0.8 \nnm Ru spacer layers, respectively. The corresponding JEX values are calculated -0.07 erg/cm2 and -0.03 erg/cm2 \nfor Ru=0.5 nm and 0.8 nm respectively. We note that the magnitude of IECE stronlgy depends not only on the \nRu interlayer thickness, but also on the magnetic and structural properties of the Ru/CoCrPt interface. In \nparticular, Peng et al. [15] in their study of CoCrPt(18 nm)/Ru(0.9 nm)/CoCrPt(8 nm) AFM system, \ndemonstrated that the additi on of a Co(1 nm) interlayer between Ru and CoPtCr, incremented J EX from 0.13 \nerg/cm2 to 0.8 erg/cm2. Generally, higher IECE is desirable for memory applications , including novel devices \nsuch as a double MTJ containing two S FM reference layers and a S FM free layer discussed in [11] which is \npredicted to switch in ps time scales. 5 \nFigure 2 . Hysteresis curves of CoCrPt(1.7)/Ru(0.8)/CoCrPt 1.3) S FM at a) 300 K, b) 50 K, and c) 2 K. The \nswitching behavior from 50-300 K includes three magnetic transitions while two transitions are observed at 2 \nK. In Figs. 2 a and 2b, the center of the o uter loops is indicated by H B. In Fig. 2c, the HB indicates the center of \nthe minor loop (red curve) associated with the switching of the thinner magnetic layer. \n An SFM has four possible magnetic configurations ( ↑↑, ↑↓, ↓↑, ↓↓), where the left and right \narrows indicate the bottom (thicker) and top (thinner) FM layers, respectively as shown in Fig. 2. The number \nof transitions found in the hysteresis loop at a given temperature depend on the IECE, the Zeeman energy, and \nthe energy barrier for magnetic reversal. From the literature, it is also evident that the magnetic field \nsweeping rate influences the magnetic switching behavior [16]. However, in this work each data point of the \nhysteresis curve is collected once the applied field has stabilized. Also, each hysteresis curve is collected once \nthe sample temperature has fully stabilized . \n Koplak et al. found for the CoFeB -Ta-CoFeB S FM three types of hysteresis loop s over the temperature \nrange studied. In the 180 K-300 K range, three magnetic transitions were observed for Type I hysteresis (↑↑-\n↑↓, ↑↓-↓↑, ↓↑-↓↓). Two transitions are present between 120 K-170 K for Type II hysteresis (↑↑-↑↓, \n↑↓-↓↓) and 2 K-110 K (↑↑-↓↑, ↓↑-↓↓) for Type III hysteresis . These transitions are observed when \nthe applied magnetic field is swept from positive to negative . The reverse transitions are encountered when \nthe field is swept from negative to positive. In the case of t he CoCrPt -Ru-CoCrPt system , we observe three \ntransitions ( ↑↑-↑↓, ↑↓-↓↑, ↓↑-↓↓) in the 50 K to 300 K range (Fig. 2a -b). Whereas t he number of \nmagnetic transitions reduces to two when the sample is cooled down to 2 K (Fig. 2c) , (↑↑-↑↓, ↑↓-↓↓) \nsimilar to the hysteresis loop measured in the 120 K-170 K range in the CoFeB -Ta-CoFeB system. The third set \nof magnetic transitions (↑↑-↓↑, ↓↑-↓↓) that are reported for the CoFeB -Ta-CoFeB S FM (Type III \nhysteresis) are not observe d for the CoCrPt -Ru-CoCrPt S FM. To understand the differences in magnetic \ntransitions between the SFM system here reported (CoCrPt -Ru-CoCrPt) and that studi ed by Koplak et al. \n(CoFeB -Ta-CoFeB) , we provide a comparative analysis based on the magnetic properties of the constituent \nlayers and on the higher IECE (4X at 300 K) provided by the Ru in our study. An important contribution of this \n 6 study is the extension of the energy balance formulation developed by Koplak et al. to SFM structures \ncomprising higher perpendicular magnetic anisotropy ferromagnets that are critical for thermally stable single \nnm scale magnetic structures. Table 1 summarizes the types of hysteresis curves observed in both the CoCrPt -\nRu-CoCrPt and the CoFeB -Ta-CoFeB SFMs as a function of temperature . \n \nIn Table 1, the hysteresis types and the associated magnetic transitions observed in the CoCrPt -Ru-CoCrPt SFM \nare shown in the second column and they are compared to the CoFeB -Ta-CoFeB SFM results reported by \nKoplak et al . [1] in the third column . The indicated magnetic transitions occur when the magnetic field is swept \nfrom positive saturation to negative saturation. The bold arrows represent the magnetic moment, m1, of the \nthicker magnetic layer. The temperature range where each type of hysteresis curve is observed are provided \nunder the heading of the different SFM structures. Three magnetic transitions are reported by Koplak et al. for \nthe CoFeB -Ta-CoFeB, while two magnetic transitions are observed for the CoCrPt -Ru-CoCrPt SFM. \nHere we analyze the magnetic transitions exhibited by the SFM with a 0.8 nm Ru spacer (Fig. 1) using \nthe energy diagram technique. This technique relies on a simple e nergy balance (Eq. 1) which contains the \nIECE (EEX), Zeeman energy (E Z), and the potential barriers E eff1 and E eff2 (corresponding to the 1.7 nm and 1.3 \nnm CoCrPt, respectively). \nEq. 1) 𝑬𝑻𝒐𝒕𝒂𝒍 =𝑬𝑬𝑿+𝑬𝒁+𝑬𝒆𝒇𝒇𝟏 +𝑬𝒆𝒇𝒇𝟐 \nThe IECE, EEX, is proportional to the surface area of the sample and can be estimated from |𝐸𝐸𝑋|=𝐻𝐵∙\n𝑚2. Here HB represents the bias field, which is measured by locating the center of the minor loop of the softer \nmagnet as previously described . In the case of the f irst type of switching shown in Fig s. 3a and 3 b, there are \nthree loops: the center field of the outer loops is HB and indicates the strength of IECE. The potential energy \nbarrier separates the two perpendicular orientations of magnetization . Notably the hysteresis curves are \nmeasured along the easy -axis, therefore, Eeff1 and E eff2 are not equal to the anisotropy energy determin ed \nfrom the hard axis hysteresis. In a hysteresis loop with three transitions (Figs. 3a and 3 b), E eff1 can be \n \n \n \n \n \n , , \n \n , \n \n , 7 estimated from the coercive field of the outer loops as 𝐻𝐶−𝑜𝑢𝑡𝑒𝑟 =𝐸𝑒𝑓𝑓 1\n2∙𝑚1. Then E eff2 can be calculated from the \ncoercive field of the inner loop expressed by 𝐻𝐶−𝑖𝑛𝑛𝑒𝑟 =𝐸𝑒𝑓𝑓 1+𝐸𝑒𝑓𝑓 2\n2(𝑚1−𝑚2). These estimat es arise from the equations \nderived by Koplak et al. describing the possible magnetic transition s. The Zeeman energy, E Z, is proportional to \nthe applied magnetic field and can be expressed as 𝐸𝑍=−(𝑚1+𝑚2)∙𝐻. \nThe second type of hysteresis curve shown in Fig. 3c has no outer loops, which are needed to estimate \nEEX. However, one can still calculate H B and thus E EX by measuring the minor loop as s hown in Fig. 2c. This is \nobtained by switching the softer, thinner m 2 magnetic layer after the SFM has been saturated [9,15] . This \nmethod of measuring the bias field advances the application of the energy balance framework to SFM \nstructures where there are no outer loops in the hysteresis curve . HB, labeled in Fig. 2c, was determined to be \n-1798 Oe. The minor loop in Fig. 2c is measured by saturating the SFM to the ↓↓ orientation , sweeping the \nmagnetic field to just beyond the ↓↓-↓↑ magnetic transition, and then saturating the SFM back to the ↓↓ \norientation. This indicates an IECE of -0.11 erg/cm2 for the S FM at 2 K. After E EX is obtained, E eff1 and E eff2 can \nbe calculated using equations 𝐻↑↑−↑↓=2|𝐸𝐸𝑋|−𝐸𝑒𝑓𝑓 2\n2𝑚2 and 𝐻↑↓−↓↓=−2|𝐸𝐸𝑋|+𝐸𝑒𝑓𝑓 1\n2𝑚1, respectively. The resulting \nenergy diagram (Fig. 3c) is consistent with the transition fields of the minor loop and the satu rated hysteresis \nloop. Eeff1 and E eff2 are plotted at each temperature in Fig. 5a. We note that , Eeff1 is larger than E eff2 until the \ntemperature is lowered to 2 K , most likely due to the changing m 1/m 2 ratio as temperature is decreased. It is \nnoted that the error associated with the E eff parameters is around ± 1x10-5 erg so the observed difference in \ntheir values is significant. A similar behavior was observed by Koplak et al . which is observed at 100 K and is \nattributed to the changing ratio of m 1 to m 2 as temperature is decreased. \n \n \nFigure 3 . Energy diagrams of the CoCrPt(1.7)/Ru(0.8)/CoCrPt(1.3) S FM at a) 300 K, b) 50 K, and c) 2 K. The solid \nlines indicate the total energy , excluding the energy barriers, while the dashed lines include the temperature -\ndependent energy barrier term, E eff. The hysteresis curves are shown in each pane with corresponding \n \n \n \n \n 8 magnetic moments on the secondary axis. The red hysteresis loop in c) displa ys the minor loop measured to \ndetermine HB. Dashed arrows indicate the energies associated with the minor loop and the corresponding \ntransitions. \nThe energy diagrams shown in Fig. 3 describe the magnetic transitions occurring for each temperature. \nSolid li nes indicate the energy of the S FM system with zero E eff, i.e. when there are no energy barriers to \novercome. The solid lines are thus , the addition of the Zeeman energy and the IECE, with a y -intercept equal \nto the IECE. The dashed lines represent the total energy of the system after the E eff1 and E eff2 are included, as \ndescribed by Eq. 1. As the magnetic field is swept, the magnetic orientation present is the one with the lowest \nenergy. Shown in Fig. 3a, the blue solid line represents the ↑↑ orientation as the field is lowered from +2 T. If \nthe energy barrier for reversal of each layer is zero, the ↑↑-↑↓ will occur at the intersection of the blue \nsolid line and the orange solid line representing the total energy of the ↑↓ orientation. At 300 K, the Eeff \nenerg ies for m 1 and m 2 are negligibly low such that the magnetic transition s occur approximately at the solid \nline intersections. \nThe potential barriers become larger as the temperature is decreased to 2 K (Fig. 5a). Since the E eff \nterms are not field -dependent, they shift the dashed lines up along the y -axis. For a magnetic transition to \noccur , the magnetic field must be changed such that the energy barrier between the solid and dashed line is \ncrossed. As seen in Fig. 3b, the ↑↑-↑↓ transition no longer occurs at the intersection of the solid blue and \norange lines, but at the point where the potential barrier of another orientation energy is crossed. As the \npotential barriers increase with lower temperature, certain tra nsitions are prohibited from occurring due to \nthe existence of lower energy states from other magnetic orientations. At 2 K (Fig. 3c ), the ↑↓-↓↑ transition \ndoes not occur as it does at 300 K and 50 K (Fig. 3a -b) since the potential barrier of the ↓↓ state is lower in \nenergy than the ↓↑ state. \nThe third set of magnetic transitions (↑↑-↓↑, ↓↑-↓↓), or Type III hysteresis, occurs when the \ncondition 𝐸𝑒𝑓𝑓 1<𝐸𝑒𝑓𝑓 2∙𝑚1\n𝑚2−2|𝐸𝐸𝑋|∙𝑚1−𝑚2\n𝑚2 is satisfied [1]. The CoCrPt -Ru-CoCrPt SFM studied does not \nmeet this requirement and does not show this set of transitions even down to 2 K. Compared to the CoFeB -Ta-\nCoFeB SFM reported by Koplak et al . [1], which has an IECE at 300 K of EEX/S = -0.01 erg/cm2, the CoCrPt -Ru-\nCoCrPt SFM has an IECE at 300 K of E EX/S = -0.04 erg/cm2. The CoFeB -Ta-CoFeB SFM and the CoCrPt -Ru-CoCrPt \nSFM have m 1/m 2 ratio s of 1.38 and 1.79, respectively . At 300 K the effective anisotropy energy barriers for \nboth systems are: E eff1/S = 4.0x10-3 erg/cm2 and E eff2/S = 2.5x10-3 erg/cm2 for CoFeB -Ta-CoFeB, E eff1/S = 1.7 x10-3 \nerg/cm2 and E eff2/S = 0.73x10-3 erg/cm2 for the CoCrPt -Ru-CoCrPt. The energy barriers for the CoCrPt -Ru-\nCoCrPt are lower for E eff1 and E eff2 by a factor of 2.3 and 3.4, respectively. This disparity in E eff causes the right 9 side of the inequa lity to be lower, which explains why the third set of magnetic transitions are absent in the \nCoCrPt -Ru-CoCrPt SFM. \nFigure 4 illustrates the dependence of the left and right side s of the inequality 𝐸𝑒𝑓𝑓 1<𝐸𝑒𝑓𝑓 2∙𝑚1\n𝑚2−\n2|𝐸𝐸𝑋|∙𝑚1−𝑚2\n𝑚2 on the IECE for both the CoCrPt and CoFeB SFM systems. The plot shows the energies \ncalculated at 100 K, since CoFeB -Ta-CoFeB shows the third set of magnetic transitions at this temperature. At a \ngiven IECE, the third set of magnetic transitions should be observed whe n the right side of the equation is \nlarger than E eff1. When plotted as a function of E EX, the right side of the inequality is a line whose slope is \ndetermined by the m 1/m 2 ratio and the intercept by the product of Eeff2 and m 1/m 2. The dependence of the \nright side of the inequality on m 1/m 2 is also illustrated in Fig. 4a -b. As m 1/m 2 approaches 1, the slope and the \ny-axis intercept of the right side of the inequality is lower ed. Since it has been shown that fast spin transfer \ntorque switching can be achiev ed with a low m 1/m 2 ratio [11], this analysis is important in understanding the \ntype of hysteresis curve s that will be present in the SFM when tailoring the ratio of magnetic moments. It is \nevident from Fig. 4a that the magnetic switching behavior of the CoCrPt -Ru-CoCrPt SFM will not exhibit the \nthird type of magnetic switching for either stronger or weaker IECE as the intercept of the right side of the \ninequality is lower than E eff1. \nAs mentioned earlier, the SFM can be used as a replacement for a single FM layer in a memory device. \nSuch devices are expected to operate successfully over a wide range of temperatu res. Therefore, it is \nimportant to predict the behavior of the SFM at any temperature. The energy diagram technique can be used \nto predict the transition fields of the SFM if the temperature dependence of E eff, EEX, and the magnetization, \nm, are known. Fig ure 5a-c show s the temperature dependence of E eff, EEX, and m, respectively. Eeff, Eeff1, and \nEeff2 are proportional to 𝑚𝑛(𝑛+1)\n2 at lower temperatures (<150 K), while at higher temperatures the potential \nenergy barriers are proportional to 𝑚𝑛 similar to the temperature dependence of the magnetic anisotropy \nobserved for other materials [17,18] . Here, m is the magnetic moment and n is the exponent of the magnetic \nanisotropy function (n=2 is typical for uniaxial anisotropy). Fig. 5a shows the fit for E eff2 based on the 𝑚𝑛(𝑛+1)\n2 \nproportionality. Good agreement is observed with the fit until around 150 K, above which there are significant \ndifference s between the meas ured E eff and the fit. At higher temperatures, Eeff2 is proportional to 𝑚𝑛 as \npredicted in [17]. Eeff1 shows a similar temperature dependence as Eeff2 and the same exponent for the \nmagnetic anisotropy function fits the data , however the data point at 50 K significantly deviates from the \ntrend. The parameters E EX and m change linearly in this temperature range . The m values are later \nextrapolated to 370 K using this linear fit. This is appropriate since this temperature is well below the Curie 10 temperature reported for Co 70Cr18Pt12 (~ 673 K [19]), around which the temperature dependence of m would \nsignificantly depart from the linear trend . It is noted in Fig. 5c that the m 1/m 2 ratio across the temperature \nrange studied is larger than the value of 1.3 expected from the nominal layer thickness of 1.7 nm and 1.3 nm. \nAs an example , at 2 K and 300 K the ratio is 1.38 and 1.7, respectively. The discrepancy may result from the \nmanner in which the magnetic moments are determined for the results presented in Fig. 5c, namely, they are \nestimated from the magnetization measurements of the SFM hysteresis loops at saturation and remanence. It \nis feasible that at remanence, contributions from the thin film multi -domain magnetic state may influence the \nmagnitude of the measured magnetization. In addition, whereas the film thicknesses and moment s of the \nconstituent layers were determined in samples grown on Ta(5 nm)/Ru(10 nm) underlayers, in the SFM, m2 is \ngrown on thin Ru , therefore its crystalline quality may differ from m 1. This could also result in difference s in \nmagnetic properties of the layers as a function of temperature. \nFitting the trends seen in Fig. 5 allows one to predict the behavior at 2 00 K and 370 K, the temperature \nextrema that a spintronic sensor or memory device could potentially expected to ope rate reliably. The energy \ndiagrams for the CoCrPt(1.7)/Ru(0.8)/CoCrPt(1.3) SFM at these two temperatures are shown in Fig. 6. The \nenergy diagram in Fig. 6a is interpolated from the parameters depicted in Fig. 5, while the energy diagram in \nFig. 5b is extr apolated from fitted parameters in Fig 5. These energy diagrams are constructed using the fitted \nparameters from Fig. 5. Both energy diagrams in Fig. 6a and 6b show magnetic transitions corresponding to \nthe type I regime [1]. The transition field for ↑↑-↑↓ (where m 1 reversal occurs) is predicted to be 1400 Oe \nand 950 Oe at 2 00 K and 370 K, respectively. In Fig. 6a the hysteresis diagram measured at 200 K is shown and \nagrees with the predicted transitions from the energy diagram. The hysteresis curve at 370 K was not be \nacquired due to errors introduced by deformation of the straw sample mounting arrangement employed for \nall other measurements as discussed in the materials and methods section . \nThe CoCrPt -Ru-CoCrPt SFM system provides significant advantages over CoFeB -Ta-CoFeB as free layers \nin MTJ structures. These derive from the lower saturation magnetization and its superior magnetic anisotropy \n[14]. The anisotropy in CoCrPt is largely determined by its magnetocrystalline anisotropy as opposed to \ninterface anisotropy for the case of CoFeB. Therefore, Co CrPt films exhibit PMA for film thicknesses up to 15 \nnm. Thus, the m 1/m 2 ratio in the films can be controlled more precisely, allowing for more tunability of the \nSFM properties. This is of particular interest for the implementation of ps magnetic switching employing SFM \nstructur es as proposed by Camsari et al. [11]. Said SFMs require larger IECE than that provided by Ta layers in \norder to achieve ps magnetic switching. Therefore, it is important to extend the energy diagram technique 11 developed by Koplak et al . to this material st ructure to understand its behavior at temperature regimes of \ninterest for practical utilization of spintronic memory devices exploiting these SFM s. \n \nFigure 4 . The left and right sides of the ine quality 𝐸𝑒𝑓𝑓 1<𝐸𝑒𝑓𝑓 2∙𝑚1\n𝑚2−2|𝐸𝐸𝑋|∙𝑚1−𝑚2\n𝑚2 plotted as a function of \nEEX for the a) CoCrPt -Ru-CoCrPt and b) CoFeB -Ta-CoFeB SFM. The black solid and dashed lines labeled “Right – \nm1/m 2” represents the right side of the inequality at different magnetic ratios. The left side of the inequality is \nrepre sented by the red curve and is labeled E eff1. The black squares shown in a) represent the observed IECE of \nthe CoCrPt -Ru-CoCrPt SFM at 100 K. \n \nFigure 5. a) Eeff1, Eeff2, and E eff (Total) plotted versus temperature. b) |E EX| plotted versus temperature. c) The \nmagnetic moments, m 1 and m 2 of the 1.7 nm and 1.3 nm thick CoCrPt layers, respectively, plotted versus \ntemperature. \n \n 12 \nFigure 6. The energy diagram for the CoCrPt(1.7)/Ru(0.8)/CoCrPt(1.3) SFM at 2 00 K and 370 K. The energy \ndiagram in a) is interpolated from the parameters de rived in Fig. 5, while the energy diagram in b) is \nextrapolated from fitted parameters in Fig 5. The t ransit ion fields are indicated by the vertical dashed lines. \nThe hysteresis curve measured at 200 K is shown in a). \nConclusions: \nMTJ devices incorporating SFM structures as free layers are expected operate reliably over a wide \ntemperature range that under extrem e conditions could range from 200 K to 370 K. Therefore, it is important \nto determine the temperature dependence of their magnetic transitions to engineer key material properties \nsuch as IECE and the energy barriers associated with magnetization reversal. In this paper the IECE of \nCoCrPt/Ru/CoCrPt SFMs has been investigated from 2 K to 300 K. Building on previous work by Koplak et al. \n[1], this work further elucidates the temperature -dependence of the potential barrier for magnetization \nreversal. The magnit ude of H B was derived from minor loop measurements, as the hysteresis curves exhibited \nno outer loops, a procedure reported also in refs. [9,15]. H B is needed to estimate IECE and details of the \nenergy diagram technique first described by Koplak et al. [1]. In this stduy, we observe two types of hysteresis \ncurves in CoCrPt/Ru/CoCrPt SFMs: one above 50 K with three subloops and the other at 2 K with two magnetic \ntransitions. Type III hysteresis, seen in the CoFeB -Ta-CoFeB SFM system in Koplak et al., is not observed in the \nCoCrPt/Ru/CoCrPt SFM, due to the large E eff1 which prevents the Eeff160 GHz) t o \nexcite the sample, and that are synchronized to the EUV probe with a controllable phase shift. The \nsecond is the optical half that ultimately focuses femtosecond -duration extreme ultraviolet (EUV) lig ht \npulses onto the sample to probe the precession of spins therein. Finally, the EUV light that reflects from \nthe sample is collected in a grating -based spectrometer. \n \nExtreme Ultraviolet Light Probes Element -Specific Magnetic Moment \n \nWe use a regenerative Ti:sapphire laser amplifier to produce 3 kHz, ~1 mJ, 35 fs pulsed light with 800 nm \nwavelength. We use this to perform high harmonic generation (HHG) by focusing it into a hollow -core \nglass capillary filled with 300 –700 Torr Neon gas . This produces a spectrum of discrete harmonics with \nphoton energies spanning 35 –70 eV , thus span ning the M -edges of most magnetic elements. Moreover, \nthese harmonics are separated by 3 eV , are pulsed at a 3 kHz rate with a time duration shorter than the \ndriving laser ( typically < 10 fs) , and have a linear polarization matching that of the driving infrared ( IR) \nlight . We use horizontally polarized light so that the beamline can be in t he plane of the optical table \neven after reflecting from the sample around Brewster’s angle for the EUV . \n Once the EUV is generated, we attenuate the driving IR using two mirrors oriented near Brewster’s angle \nfor the IR , and further extinguish the IR by absorbing it in thin metal foil(s) (usually 0.5 μm Al). Then, the \nEUV passes through a variable -diameter aperture, and is focused onto the sample using a 6° angle of \nincidence ( AOI), grazing -incidence toroidal mirror. This mirror is placed 2 m from the glass capillary and \nhas a focal length of 750 mm, resulti ng in a spot size of approximately 40 μm diameter [the width of the \ncenter conductor of the coplanar waveguide sample s, Fig 2 (c)]. Since we are using the T -MOKE \ngeometry [see Fig. 2 (b)], the sample is oriented near Brewster’s angle, specifically at 50° from grazing \nsince this angle maximizes signal from many of our samples (see Fig. 4). We can energize a vertically -\noriented coil magnet for static T -MOKE measurements or a horizontally -oriented electromagnet for our \nXFM R studies . Both magnetic fields are parallel to the sample plane. By reversing the direction of the \nmagnetic field or (for XFMR) by toggling the RF on/off with a fixed magnetic field, we can observe the \nelement -specific magnetic state of the sample as the variation of the spectral peak in tensity at the \nabsorption edge of each element. In both cases , the projection of the magnetic moment onto the \nvertical direction is the detected magnetic signal — for XFMR, th e signal will oscillate sinusoidally with \nthe spins’ precession . Finally, the reflected EUV strikes a 400 mm toroid in a 4f imaging configuration \n(magnification of 1) and a 500 grooves/mm blazed grating (period: 2 μm) oriented in the conical \ngeometry, and is collected on a scientific EUV CCD detector. This results in a spectrum with separated \npeaks , shown in Fig. 2 (a). \n \nGeneration of synchronized >60 GHz microwaves \n \nFigure 2. Probing Element -Specific Magnetization : (a) Example EUV spectra after reflecting from a gold \nsample (relatively flat spectral response) and from a magnetic film (permalloy). The dips in reflectivity \nindicate the absorption edges of the elements present (Fe and Ni). The red and blue regions i ndicate the \nposition and width of the Fe and Ni absorption edges. (b) Picture of the sample chamber, which is \nconfigured in the transverse magneto -optic Kerr effect (T -MOKE) geometry. A coil magnet placed a bove \nour sample (indicated, but not shown) enables us to measure the static magnetic signal from our \nsample. The higher -field electromagnet has an in -plane, horizontal axis; during XFMR, the spins precess \nabout the horizontal axis and the projection of th e resulting magnetic moment onto the vertical \ndirection is the measured signal. (c) Picture of a sample. Each sample is a lithographically patterned 50 \nΩ coplanar waveguide with magnetic layer(s) patterned on top of the center conductor. The center \ncond uctor is 40 μm wide, and matches the size of the EUV focus. An additional, much larger region of \nthe same magnetic layer is patterned far from the waveguide for verification of static signals. \nThe pulse width of the EUV is <30 fs, which sets an intrinsic frequency limitation well above 10 THz. \nTherefore, the practical frequency limitation for XFMR with this instrument is set by the signal -to-noise \nratio (SNR, which is sample -dependent and decreases at higher frequency due to smaller precession \ncone angle s) and the timin g jitter between the microwave excitation and the EUV probe , which has a \nstandard deviation of roughly 1.3 ps, as seen in Fig. 3 . \n \nTo maximize the SNR, we adopt an on/off measurement scheme, in which every other measured \nspectrum is made without the microwave signal applied. By subtracting the spectrum with the \nmicrowaves off from that with the microwaves on, the effect of longer -timescale fluctuations of the EUV \nintensity can be minimized . Since our exposure times were 5 seconds, the effects of fluctuations on this \ntimescale or shorter are not suppressed. We partially correct these shorter -term instabilities by \nFigure 3. Using YIG -Tuned filters as a broadband phase shifter : (a,b) Timing jitter of 17 GHz and 51 \nGHz microwaves, measured using a 40 GHz sampling oscilloscope. (c) Spectrum analyzer trace \nshowing that the neighboring harmonics are suppressed by >30 dB, resulting in a pure, narrow -\nlinewidth, sine wave excitation. (d) Imparted phase shift as a function of YIG control current, \nshowing that t he same control current provides greater phase shift at higher frequencies. \nassuming that they are perfectly correlated across all the harmonic peaks (including those with no \nmagnetic signal) and dividing all peaks by the frame -to-frame fluctuations of the peak intensitie s far \naway from the magnetic absorption edges . This is only approximately correct, but consistently improves \nour data quality. \n \nTo minimize the timing jitter between the microwave excitation and the EUV probe , we generate the \nmicrowaves directly from the IR laser. In particular, w e use a similar approach to that taken at \nsynchrotron facilities, wherein an RF frequency comb generator (FCG) produces harmonics of the master \nclock that times the el ectron bunches.9,19 In this case, we adopt the Ti -sapphire oscillator itself as the \nmaster clock (81.6 MHz). A portion of the oscillator beam is directed onto a photodiode, and the \nresulting electrical signal drive s an RF frequency comb generator that produces harmonics up to at least \n18 GHz (> 220th harmonic). Since the oscillator is simultaneously used to seed the regenerative amplifier , \nthe EUV pulses are intrinsically synchronized to the microwave radiation . \n \nAn upper bound of the short -term phase jitter can be made by measuring the microwave signal with a \nsampling oscilloscope that is triggered on a fast photodiode (rise time: 30 ps) that views the 3 kHz laser \nthat drives the HHG process. Shown in Fig. 3 (b) is a measurement of 5 1 GHz microwaves made in this \nway, which shows a timing jitter of 1. 4 ps. Due to the limited bandwidth of the available microwave \ncomponents, we used different configurations for 8.5 GHz, 17 GHz, and 62 GHz frequencies (see Fig. S2) . \nImportantly, the timing jitter was roughly the same (<1. 4 ps) for each configuration, even when we used \nadditional components like frequency multipliers. The actual jitter may be lower than 1. 4 ps, owing to \nthe fact that smaller values are beyond the measurement bandwidth of the oscilloscopes. Furthermore, \nthe tim ing jitter may be further improved by use of a high -speed photodiode as input to the FCG, instead \nof the current one which has 1 ns rise time. Considering this upper limit of timing jitter, there is no \nfundamental limitation to achieve measurement frequencies of 100 GHz or higher. \n \nWhile we need to excite FMR with a single frequency, t he FCG outputs a large series of spectral lines that \nspan several hundred harmonics of the fundamental 82 MHz repetition rate of the oscillator . As shown \nin the simplified schematic diagram in Fig. 1 (b) and in the more complete diagrams in Appendix B, we \nuse a pair of YIG filters as a narrow -band filter to isolate a single frequency or “tooth” of the frequency \ncomb [suppressing other frequencies by more than 30 dB, as shown in Fig. 3(b)]. In addition to filtering \nout undesired frequencies , the YIG filters shift the phase of the signal as they are detuned from their \nresonance; we exploit this property for use as a self-contained broadband microwave phase shifter that \nrequires no additional components (see Appendix A). Of course, the total phase range is limited by the \nwidth of the YIG resonance. Using two YIGs in tandem ( instead of one ) gives a larger range of access ible \nphase shifts while keeping the amplitude relatively flat. F or the frequencies investigated here, using two \nYIG filters enables the phase to be continuously shifted over a range of at least 270 degrees (8.5 GHz), \n540 degrees (17 GHz), and with the range generally increasing in proportion to frequency [see Fig. 3 (d)]. \n \nWith our current setup , we are able to vary the base frequency of the FCG from 8.5 GHz to 12 GHz. This \nlimitation is simply set by the bandwidth of the microwave components (including YIG filters, amplifiers, \ncirculators, high -speed switch, etc.) that we had available on hand ; however, standard , commercially \navailable products can be used for other /wider bands , extending both to lower and higher frequency \nranges. In our current setup, h igher frequencies are achieved by inserting frequency multipliers (2x and 3x) that enable us to extend the output frequencies to values exceeding 60 GHz (see Appendix B). The \nadvantage to this approach is that most of the circuit remains unchanged (i.e., operates at its base \nbandwidth) since the frequency multiplier is inserted just before the sample , requiring only one higher \nfrequency amplifier and components after the multiplier . Figure 3 (a, b) shows an example of a 17 GHz \nsignal generated by frequency doubling 8.5 GHz, and a 51 GHz signal generated by using both a \nfrequency doubler and a frequency tripler. A final note is that , by introducing a circulator and a lock -in \namplifier [shown in Fig. 1 (b)], we are able to perform in situ inductive FMR measurements on the \nsample . This not only provides information about the optimal fields to use during the XFMR \nmeasurement, but also allows us to moni tor the sample for an y possible damage , heating, or other \nchanges during the measurement. \n \nXFMR Procedure \n \nAfter we have generated femtosecond -duration EUV light in the appropriate range of photon energies \n(45–70 eV) and generated microwaves that are phase -synchronized to those EUV pulses, we can perform \nXFMR using the following steps . We first set the excitation microwave frequency (e.g., to 17 GHz) by \ncoarsely tuning the YIG filter current s to select a tooth of the RF frequency comb. Fine -tuning the \ncurrent s about this point will control the phase shift of the microwaves. To quantify this in situ, we use a \nsampling oscilloscope, which is triggered on the 800 nm laser that drives HHG, to measure the \nmicrowaves, and sweep the YIG voltages independently to determine the phase shift per volt (approx. \n600 deg/ mA at 8.5 GHz) and the accessible range of phase shifts . \n \nAfter setting the frequency and calibrating the phase -shifting YIG control current s, we use our in situ \ninductive FMR setup to determine the magnetic fields to perform XFMR (i.e., resonance field and \nlinewidth) , as well as the proper microwave power to maximize our signal but suffer no effects from \nheating and ensure that we stay within the linear regime of the magnetization dynamics . Th e inductive \nFMR trace will show a linear combination of the real and imaginary parts of the microwave reflected \nfrom the sample. It is easiest to interpret the imaginary part , which we achieve by use of RF phase \nshifting components. \n \nOnce we know the fields over which to perform XFMR, we send EUV light onto the sample. We can then \nmeasure and optimize ou r system using a static T -MOKE spectrum obtained by energizing a small coil \nmagnet above our sample. By switching the polarity of the current, the magnetic state of the sample will \nreverse and this will write itself onto the harmonic spectrum; in other words , the harmonic peaks at the \nphoton energies within the Fe M -edge will change intensity proportionally to the magnetic state of the \nFe. By measuring the distance between the grating and the camera and then counting the number of \npixels between the specular beam and the harmonic peaks of interest, we can estimate the photon \nenergy of each harmonic peak and determine which element is m agnetized by using the grating \nequation ( 𝜆=𝑇sin(𝜃)/𝑞, where λ is the wavelength, T is the period of the grating, θ is the diffraction \nangle, and q is the diffraction order ). To improve the SNR o f our static T -MOKE measurements, we \naverage several frames ( typically 20 frames per magnetic field polarity ) and generally use an exposure \ntime of 5 seconds for each frame while 4x4 binning. To avoid condensation issues on the camera sensor \ndue to operating in only high vacuum conditions , we do not cool the sensor. \n Finally, if we fix the applied magnetic field and \nthe YIG currents , then each EUV pulse will \nimpinge on the sample at a consistent phase \nof the applied microwaves, and therefore at \nthe same precessional phase of the spins \nwithin the sample. We again use 5 sec ond \nexposure times for each frame and average \nover many frames ( typically 100 with \nmicrowaves applied and 100 without). A \ndifference spectrum is obtained by recording \nalternating spectra with the microwaves \nturned on/off and subtracting them from \neach other . As a result, the difference \nspectrum is proportional to the projection of \nthe magnetization on to the vertical axis at a \nparticular phase in the precession ; as the \nphase is shifted through 360 °, this measure s a \nsinusoidal response. This process can be \nrepeated at a series of magnetic fields to map \nout the full , element -resolved, field -swept \nferromagnetic resonanc e peak . Notably, the \nprocess of recording spectra with the \nmicrowaves alternately on and off mitigates \nthe effects of long -term intensity drift of the \nEUV source. \n \nSample Fabrication \n \nSample s were fabricated using optical \nlithography combined with lift-off processes \non thermally oxidized Si wafers. The first step \nwas to pattern the waveguide structure. This \nconsist s of a coplanar waveguide (CPW) \nstructure that is designed to mate with a \ncommercial high -frequency end-launcher on \none end. This connection is designed to \nmaintain a 50 Ohm impedance, minimizing \nreflections or losses across that transition \nto/from the end launcher . The CPW then tapers down to a center conductor width of 40 μm in most \ncases, but other widths are also used . To form the CPW, 5 nm Ti and 100 nm of Au was evaporated and \nlifted off. A second lithograph y step was used to define the magnetic lay er of interest. This was \npatterned on top of the narrow section of the CPW , as shown in Figure 2(c). The magnetic material was \nthen deposited using magnetron sputtering . In all cases, a 3 nm layer of TaO x was first deposited prior to \ndepositing the magnetic layer in order to prevent electrical contact between the magnetic layer and the \nFigure 4. Samples and selection of measurement \nangle: The three magnetic sample structures are \nshown schematically in (a -c). On the left are the \nstructures that were deposited on top of the gold \ncenter conductor, and on the right are the simulated \nmagnetic signal vs. photon energy and incidence \nangle for e ach sample. We selected 50 ° from grazing \nfor its sensitivity to our permalloy sample. The other \nsamples would have had higher signal at steeper \nincidence angles. \nCPW. This also prevent s any spin -\npumping losses when the magnetic layer \nis driven to FMR . After lift -off of the \nmagnetic layer, the wafer was diced using \na diamond saw. \n \nIn this study, we focus on three magnetic \nsamples , whose structures are \nsummarized in Figure 4. First, we \nperformed XFMR on permalloy at 8.5 \nGHz. This sample could not be used to \ndemonstrate XFMR at frequencies above \n11 GHz due to the magnetic field \nlimitation of our current setup (max μ0 H = \n180 mT). We instead focus on a Co25Fe75 \nsample to achieve higher -frequency XFMR \nmeasurements at 17 GHz . The increased \nsaturation magnetization of µ0Ms = 2.4 T \nincreases the FMR frequency to more \nthan 17 GHz for the magnetic fields that \nwe have available.56 Finally, to show the \nability to measure independ ently the \ndynamics of different layers with in a \nmultilayer sample, we measured a \nNi/TaO x/Fe multilayer . \n \nXFMR Results: Permalloy at 8.5 GHz \n \nAs a first demonstration , we measure d \nthe XFMR response of a 10 nm thick \nNi80Fe20 (permalloy) sample. The \nindividual spectra at the microwave phase \ngiving the maximum magnetic signal are \nshown in Figure 5 (a). Two spectra are \nshown: o ne each with the microwaves \nturned on and turned off. There is a small \nbut measurable difference between the \ntwo spectra that measures spin \nprecession within the sample; this signal \nis significantly enhanced at the expected photon energies corresponding to the M -edges of Ni and Fe , \nand varies sinusoidally with the phase of the applied microwave excitation . Th e difference between the \ntwo curves is related to the magnetic asymmetry , which is calculated as (Ion - Ioff)/(Ion + Ioff), or the \ndifference between the intensiti es of the spectra when the rf field is on Ion and when it is off, Ioff divided \nby the sum of the intensities . The magnetic asymmetry is show n in the bottom half of Figure 5(a). For \nFigure 5. Element -resolved XFMR on permalloy, 8.5 GHz: \n(a) The top plot shows the raw spectra, and the bottom \nplot shows an example XFMR asymmetry spectrum. \nThere are peaks in the magnetic asymmetry at the M -\nedges of Ni and Fe, indicating the amplitude of the \ndynamic response. (b,c) show the measured spin \nprecession vs. time for Fe and Ni. (d,e) show the XFMR \ntraces of the individual elements. Overlaid are the \ninductive FMR measured in situ and the simultaneous fit \nof amplitude and phase. \nthe sake of visualization, we suppress the noise in the asymmetry spectra in the region s between the \nharmonics where there is little to no signal ; we do so by adding a large value to the denominator when \nthe intensity is below a threshold that is set to a value just above the minimum counts in between the \nharmonics . This minimizes the effect of amplifying noise in the spectral region where there is little to no \nsignal and/or prevent a divide by zero . It is important to note that this is not used in the quantitative \nfitting of the spectra, but solely to improve visualization . \n \nThe amplitude of the asymmetry should vary sinusoidally at the microwave driving frequency . To \ncapture the amplitude and phase of the response for each element , the magnetic spectra are recorded \nat several values of the rf phase. The amplitudes at the Fe and Ni M -edges are plotted as a function of \nthe phase in Figures 5(b) and 5(c), respectively. A sinusoid is then fit to the data , using the frequency \nmeasured by a spectrum analyzer during the experiment . This process is repeated for all values of the \napplied bias field — a few examples are also shown in Figures 5(b) and 5(c). It is clear that , as expected, \nboth the amplitude and phase of these sinusoidal behaviors change as the magnetic field is also \nchanged. Figure 5(d) and 5(e) are element -resolved XFMR plots. That is to say, they show the amplitude \n(black circle) and phase (blue open tria ngle) obtained from the sinusoidal fits of the Fe and Ni peaks data \nfor several values of the applied magnetic bias field. The amplitude A and phase φ are simultaneously fit \nto the following Lorentzian and sigmoidal arctangent function s, respectively : \n \n𝐴=𝑦0+(2𝐴0\n𝜋)𝑤\n√4(𝐻−𝐻0)2+𝑤2 \n \n𝜑=𝜑0+180°\n𝜋(arctan(𝐻−𝐻0\n𝑤)+𝜋\n2) \n \nwhere, y0 and φ0 are offsets, A0 is the amplitude of the resonance, w is the linewidth, H is the magnetic \nfield, and H0 is the resonance field . The simultaneous fit is indicated in the figure as solid lines . The \nidentical response between the Fe and Ni is expected based on the strong exchange coupling and \nstandard models for FMR. This provides rudimentary validation of the approach . Furthermore, t he \nquality of the simultaneous fit highlights the agreement between the measured and expected behavior \nas the external bias field is swept th rough the ferromagnetic resonance. To confirm the measurement of \nthe FMR, we overlay as a red line the field -swept FMR spectrum that we obtain via the in situ inductive \nmethod by measuring the amplitude of the reflected microwaves from the CPW sample structure. Good \nagreement is seen between the XFMR and inductive FMR methods. The small difference in linewidth \nbetween the inductive FMR and XFMR could be due to the fact that XFMR is probing a localized area of \nthe sample, whereas the inductive approach is probing the entire sample which will likely have more \ninhomogeneity. This is especially true given the fact that the current sample was fabricated in a lift -off \nprocess on a thick Au layer. This produces edge regions with both line edge roughness and thickness \nvariation , coupled wit h the fact that the Au underlayer itself induces additio nal film roughness. Finally, \nwe note that since our EUV beam is focused in a roughly -Gaussian spot with a peak intensity near the \ncenter of the waveguide, o ur EUV measurements intrinsically probe the center region of the waveguide \nmore than the edges . \n \nXFMR Results: Co 25Fe75 at 17 GHz \n \nThe second sample that we studied is Co25Fe75; due to its high M s, we were able to measure XFMR at 17 \nGHz even with magnetic fields as low as µ0H ≈ 150 mT. Notably, at 50° from grazing , our setup has a low \ncontrast for the magnetic signal at the Co M-edge as shown in Figure 4(b). This limitation is solely due to \nthe current incidence angle of our system , and can be overcome by changing this angle to 52°. \nRegardless, we are still able to measure the dynamics of each element , as shown in Figure 6 (a) . Broadly \nspeaking, the Co and Fe behave the same, as expected; however, without more careful data analysis, \nquantitative XFMR traces cannot be shown for Co and Fe independently (e.g., to compare phase shifts) \ndue to the proximity of the Co (57 –62 eV) and Fe (50 –55 eV) M-edges . \n \nThe inductive FMR trace shown in Figure 6(c) shows a double -resonance , which we suspect is due to \ninhomogeneity in the sample. This material is known to have large inhomogeneity when deposited on \nrough surfaces or when non -ideal seed and capping layers are used.57 The XFMR trace also shows this \ndouble peak structure , and the locations of the two resonances agree in both methods . As expected, \nthe phase in the XFMR is not a clean arctan due to superposition of the two resonances. The relative \namplitudes of the two peaks are not the same in the XFMR and inductive FMR , which is likely due to the \nfact that the EUV probes only a small region (~60 μm diameter), whereas inductive FMR probes the \nentire structure. We do not think this comes from uncertainty in our XFMR data; i n addition to \nestimating the error bars from the quality of the sinusoidal fits (i.e., the variance in the data), we tested \nthe repeatability of the XFMR measurement by measuring at 150 mT both at the beginning and end of \nthe measurement [included in Fig. 6(c)] . Within their respective error bars, these measurements at 150 \nmT agree and demonstrate the stability and robustness of the measurement system. \nFigure 6. High -Bandwidth XFMR on Co 25Fe75, 17 GHz: (a) The top plot shows the raw spectra, and the \nbottom plot shows an example XFMR magnetic asymmetry. The Co and Fe absorption edges are \nclose to each other, so the dynamics are more difficult to separate quantitatively. (b) The spin \nprecession of the Fe atoms. (c) The XFMR trace of the Fe, with the inductive FMR overlaid. \n \nXFMR Results: Ni/Fe Multilayer at 8.5 GHz, showing element -specific dynamics \n \nThe third and final sample in this study is a multilayer sample with Ni and Fe layers that are separated by \nan insulating 3 nm TaO x layer. The magnetic isolation of the Fe and Ni layers should yield independent \ndynamics of the layer except for some potential dipolar coupling. Figure 7(b) shows the in situ inductive \nFMR response of the sample taken at 8.5 GHz , where in two peaks are observed. The high amplitude and \nnarrow peak near 40 mT is presumably due to the Fe layer since Fe has higher M s and lower damping \nrelative to Ni.58 The broad and low amplitude peak near 120 mT is likely from the Ni, by the same \nargument. Indeed, XFMR measurements at 40 mT reveal that the precession amplitude of the Fe layer is \nsignificantly higher than that of the Ni. Given the broad linewidth of the Ni peak and the potential for \ndipolar coupling between the layers, it is not surprising to see a non -zero precession angle for Ni as well. \nRepeating this measurement at μ0 H ~125 mT shows the opposite behavior , where the precession \namplitude of the Ni signal is significantly more than that of the Fe. Furthermore, in both cases there is a \nnotic eable phase shift between the precession of the two elements . Such a phase shift is to be expected \nsince the resonance fields (and therefore phase shift through the resonance) are at different locations. \nAny dipolar coupling would also induce a relative phase shift. Measurements taken at more fields across \nthe resonance s would in principle enable us to determine the amount and type of coupling between the \nFigure 7. Element -Resolved Dynamics in a Magnetic Multilayer, 8.5 GHz: (a) The top plot shows the \nraw spectra for static T -MOKE. The second row shows the magnetic asymmetry from the static \nspectrum, which shows contributions at both the Ni and Fe edges. The bottom two plots are \nrepresentative XFMR magnetic asymmetry spectr a taken at the Ni and Fe resonant fields. The fact that \nthere is no response of the Fe atoms at the Ni resonance field highlights the ability of XFMR to \nmeasure element -resolved dynamic s. (b) We measured XFMR at the Fe resonance and at the Ni \nresonance, and the measured element -resolved precession is shown in the rightmost plots. \nlayers. Measuring such phase shifts in structures where the coupling is more complicated (e.g., from \nspin currents or weak exchange) allows quantification of such interactions . This highlights the ability to \nseparate dynamics within different layers or sublattices in scientifically interesting or industry -relevant \nsamples . \n \nDiscussion and Outlook \n \nThis work shows that high -bandwidth , element -specific measurements of dynamic magnetic phenomena \ncan be achieved in a laboratory. Despite the fact that no reference beam was used to correct for \nintensity flu ctuations and drift, small deviations of the magnetization with cone angles below 5 degree s \nwere easily observed at high -frequency. Going forward, many straightforward improvements can be \nmade to increase the sensitivity and SNR , allowing more subtle phenomena to be observed along with \nsignificantly smaller cones angles and allow for more field steps across the resonance . The latter will be \nparticularly important as measurements are pushed to higher frequencies. Here, we discuss several \nimprovements that can be made to greatly improve performance . \n \nFirst, a reference beam that provides reference spectr a of the EUV light incident on the sample can be \nused to normalize and divide out intensity fluctuations on both short and long timescales . The absence \nof this in the current setup represent s one of our biggest barriers to achiev ing higher SNR. The frame -to-\nframe RMS intensity is approximately 5% for 5 sec exposures, and there is also similar drift on longer \ntime scales. Furthermore, the fluctuations of each harmonic are not necessarily correlated (though our \nanalysis assumes that they are) . Such normalization is critical at synchrotrons since the beam intensity \nalso varies significantly over time. Recently, the use of a reference beam in HHG systems used to \nmeasure spin dynamics was shown to increase SNR by almost an order of magnitude .59–61 \n \nSecond, reflection -geometry measurements as shown here can be significantly improved by \nincorporating measurements at multiple angles .39,55 This can be seen in Figure 4, which shows a \ncalculation of the T-MOKE contrast for the samples used in this study. This calculation was performed \nusing the scattering matrix approach described in Ref. [62]. While the 50° angle of incidence (measured \nfrom grazing incidence) used here is near the maximum -contrast point for the Ni 80Fe20 sample, we would \nachieve better SNR and better -separated Ni and Fe peaks in the Co25Fe75 and Ni/Fe multilayer sample s if \nthe scattering angle was changed to 52 degrees. Incorporating the ability to change the incidence angle \neasily will enable us to optimize the signal for general samples, as well as improve the depth -sensitivity \nof our measur ement. \n \nThird, while we have shown adequate synchronization of the EUV pulses to 62 GHz microwave fields, the \ntiming jitter we measure supports extension to significantly higher frequencies by using commercially \navailable components. Furthermore, it may be possible to improve the timing jitter by replacing the \nphotodiode that is used for the input to the frequency comb generator . We found that using a sharper \nrise time on the input to the frequency comb generator reduced the timing jitter. Since t he current \nphot odiode still has a relatively long rise time (low bandwidth) of 1 ns , which depending on the \nelectronic noise present can produce additional timing jitter , we anticipate that using a photodiode with \na shorter rise time will further improve the timing jitter . Also, i n addition to using multipliers , \nsynchronized 100+ GHz tones with higher signal -to-noise ratio c an be generated by driving an electro optic modulator with a 10 or 20 GHz harmonic of the 80 MHz laser repletion rate .63 We further \nnote that our current Ta -sapphire oscillator is passive and lacks any active frequency stabilization. \nFemtosecond -level timing jitter between optical and electronic signals is accessible with techniques that \nemploy stabilization of the underlying Ti:sapphire frequency comb mode spacing and carrier -envelope \noffset frequency .63–65 \n \nFinally, as already mentioned, this technique will be extended to include an imaging modality whereby \nthe coherence of the EUV will be exploited .51,53 Lensless imaging techniques such as coherent diffractive \nimaging, ptychography, and holography have been used with EUV and x -ray light to image \nnanostructures and magnetic textures with resolution smaller than the illumination spot size and \napproaching the diffraction limit .66,67 Furthermore, these techniques provide quantitative measurement \nof the phase shift imparted by the sample, which provides a dramatic improvement in contrast to certain \nsample features , such as topography and oxidation state .55 Combining such a technique with the XFMR \ninstrument developed in this paper will enable us to perform dynamic, in situ measurements of \nfunctioning devices and elucidate the underlying physics. \n \nSummary \n \nWe have demonstrated the ability to bring element -specific, x -ray detected ferromagnetic resonance \nspectroscopy (XFMR) to a laboratory setting. We did so by combining a high -harmonic generation (HHG) \nlight source , which generates 35 –70 eV light, with an RF frequency comb generator, and were able to \ngenerate phase -stable microwaves (timing jitter < 1. 4 ps) above 60 GHz that we can use for XFMR \nmeasurements at the M -edge of most magnetic elements. The instrument can perform measurements \nin transmission or r eflection; we showed reflection -mode measurements here that use the transverse \nmagneto -optic Kerr effect (T -MOKE) geometry for magnetic contrast even on samples with opaque \nsubstrates (e.g., thick silicon). We used our instrument to perform high -frequency XFMR measurements \non three samples: permalloy (8.5 GHz), Co 25Fe75 (17 GHz), and a Ni/Fe multilayer sample that showed \ndifferent dynamics in the Ni and Fe layers (8.5 GHz) . \n \nThis system provides the capability in the near future to measure element -, layer -, or sublattice -specific \ndynamics in industry -relevant and scientifically interesting magnetic thin films. Moreover, by introducing \nvariable -angle measurements into this system, we will enhance the SNR and depth -sensitivity of the \ntechnique to be able to monitor precisely the spin transport and accumulation across interfaces. Finally , \nthe spatial coherence of the source, combined with the ability to synchronize electrical pulses to the EUV \nprobe, lays the foundation for performing these measurements in -operando on individual or arrays of \nfunctional spintronic devices. \n \nAppendix A: Tuning the YIG Pair \nWe are able to tune the pair of YIG filters by using a pair of part-per-million (PPM )-stabl e current sources \n(one per YIG). In particular, we drive the YIGs with roughly 200 mA, and changing the current by 1 µA \nshifts the phase by roughly a degree at 8.5 GHz [shown both in Fig. 3(d) and Fig. S(1a)] . One YIG alone \ndoes not provide 360° phase shift at 8.5 GHz (or below) before the transmitted amplitude starts to \ndecline. By using two YIGs and sweeping the currents together, we are able to shift the microwave phase over a full cycle while maintaining a relatively constant microwave amplitude. Fig. S1 shows the \namplitude and phase of the transmitted microwaves as the YIG filters are tuned together (vertical) and \napart (horizontal). \n \n \nAppendix B: Complete RF Generator Schematic \n \nFigure S2 shows the full schematic for the microwave generator that was used to generate 8.5–62 GHz. \nWe use frequency multipliers to achieve frequencies above 13 GHz. Above 54 GHz , we encounter \nsignificant losses in our system since we exceed the bandwidth of several components , as well as the 40 \nGHz sampling oscilloscope. Despite this fact, we are still able to evaluate the timing stability and test the \nsynchronization of microwaves to the EUV pulses up to 62 GHz. Figure S2 shows that we can still \nmeasure the synchronization our EUV pulses up to frequencies of 62 GHz, albeit with increased noise \ndue to microwave losses. Additionally , the final bandpass filter fails to suppress the 4 1 GHz harmonic \nalso output by the tripler when generating 62 GHz microwaves, leading to some beating of the signal. \nHowever, these are not fundamental limitations since higher bandwidth components are commercially \navailable. \nFigure S1. Tuning the YIG filter pair: (a) The amplitude of the transmitted microwaves is roughly \nmaintained over the passband of the filters. The dashed purple arrow shows an example range of \napproximately 360 ° phase shift at 8.5 GHz. (b) The phase is shifted if the YIG filter currents are \nchanged together. \n \nAcknowledgments \nThe authors are grateful to Henry Kapteyn, Margaret Murnane , and Tom Silva for valuable discussions \nand advice , and to Dmitriy Zusin and Christian Gentry for developing the multilayer reflectivity \ncalculating code . MT and JW ackno wledge funding support from the National Research Council (NRC) \nFigure S2. Complete RF Generation Circuit Schematic for up to 62 GHz: (a) The full circuit schematic for \ngenerating synchronized microwaves is shown. The frequency -doubling and frequency -tripling \nelectronics are omitted when generating lower frequencies. (b -d) Traces measured with our 40 GHz \nbandwidth sampling scope when the circuit is generating 40 GHz, 57 GHz, and 62 GHz microwaves, \nrespectively. In (b), the trigger pulse (measured by a high -speed, real -time oscilloscope) is overlaid. \nNote that the be ating at 62 GHz is due to exceeding the rated bandwidth of multiple components, most \nnotably the final bandpass filter passes 41 GHz in addition to the desired 62 GHz. \nPost -Doctoral Fellowship program. HTN acknowledges support through the NIST cooperative agreement \n70NANB18H006 with the University of Colorado Boulder . \n \nReferences \n \n1 D. Edelstein, M. Rizzolo, D. Sil, A. Dutta, J. DeBrosse, M. Wordeman, A. Arceo, I. C. Chu, J. Demarest, E. \nR. J. Edwards, E. R. Evarts, J. Fullam, A. Gasasira, G. Hu, M. Iwatake, R. Johnson, V. Katragadda, T. Levin, J. \nLi, Y . Liu, C. Long, T. Maffitt, S. McDermott, S. Mehta, V. Mehta, D. Metzler, J. Morillo, Y . Nakamura, S. \nNguyen, P . Nieves, V. Pai, R. Patlolla, R. Pujari, R. Southwick, T. Standaert, O. van der Straten, H. Wu, C. . -\nC. Yang, D. Houssameddine, J. M. Slaughter, and D. C. Worledge, in 2020 I EEE International Electron \nDevices Meeting (IEDM) (2020), p. 11.5.1 -11.5.4. \n2 D.C. Worledge, in 2022 IEEE International Memory Workshop (IMW) (2022), pp. 1 –4. \n3 A.V. Chumak, P . Kabos, M. Wu, C. Abert, C. Adelmann, A.O. Adeyeye, J. Åkerman, F.G. Aliev, A. Anane, A. \nAwad, C.H. Back, A. Barman, G.E.W. Bauer, M. Becherer, E.N. Beginin, V.A.S.V. Bittencourt, Y .M. Blanter, P . \nBortolotti, I. Boventer, D.A. Bozhko, S.A. Bunyaev, J.J. Carmiggelt, R.R. Cheenikundil, F. Ciubotaru, S. \nCotofana, G. Csaba, O.V. Dobrovolskiy, C. Dubs, M. Elyasi, K.G. Fripp, H. Fulara, I.A. Golovchanskiy, C. \nGonzalez -Ballestero, P . Graczyk, D. Grundler, P . Gruszecki, G. Gubbiotti, K. Guslienko, A . Haldar, S. \nHamdioui, R. Hertel, B. Hillebrands, T. Hioki, A. Houshang, C. -M. Hu, H. Huebl, M. Huth, E. Iacocca, M.B. \nJungfleisch, G.N. Kakazei, A. Khitun, R. Khymyn, T. Kikkawa, M. Kläui, O. Klein, J.W. Kłos, S. Knauer, S. \nKoraltan, M. Kostylev, M. Krawc zyk, I.N. Krivorotov, V.V. Kruglyak, D. Lachance -Quirion, S. Ladak, R. \nLebrun, Y . Li, M. Lindner, R. Macêdo, S. Mayr, G.A. Melkov, S. Mieszczak, Y . Nakamura, H.T. Nembach, A.A. \nNikitin, S.A. Nikitov, V. Novosad, J.A. Otálora, Y . Otani, A. Papp, B. Pigeau, P . Pirro, W. Porod, F. Porrati, H. \nQin, B. Rana, T. Reimann, F. Riente, O. Romero -Isart, A. Ross, A.V. Sadovnikov, A.R. Safin, E. Saitoh, G. \nSchmidt, H. Schultheiss, K. Schultheiss, A.A. Serga, S. Sharma, J.M. Shaw, D. Suess, O. Surzhenko, K. Szulc, \nT. Tan iguchi, M. Urbánek, K. Usami, A.B. Ustinov, T. van der Sar, S. van Dijken, V.I. Vasyuchka, R. Verba, \nS.V. Kusminskiy, Q. Wang, M. Weides, M. Weiler, S. Wintz, S.P . Wolski, and X. Zhang, “Advances in \nMagnetics Roadmap on Spin -Wave Computing,” IEEE Transacti ons on Magnetics 58(6), 1 –72 (2022). \n4 Q. Shao, P . Li, L. Liu, H. Yang, S. Fukami, A. Razavi, H. Wu, K. Wang, F. Freimuth, Y . Mokrousov, M.D. \nStiles, S. Emori, A. Hoffmann, J. Åkerman, K. Roy, J. -P . Wang, S. -H. Yang, K. Garello, and W. Zhang, \n“Roadmap of Spin –Orbit Torques,” IEEE Transactions on Magnetics 57(7), 1 –39 (2021). \n5 A.V. Khvalkovskiy, D. Apalkov, S. Watts, R. Chepulskii, R.S. Beach, A. Ong, X. Tang, A. Driskill -Smith, W.H. \nButler, P .B. Visscher, D. Lottis, E. Chen, V. Nikitin, and M. Krounbi, “Basic principles of STT -MRAM cell \noperation in memory arrays,” J. Phys. D: Appl. Phys. 46(7), 074001 (2013). \n6 S.S. Kalarickal, P . Krivosik, M. Wu, C.E. Patton, M.L. Schneider, P . Kabos, T.J. Silva, and J.P . Nibarger, \n“Ferromagnetic resonance linewidth in metallic thin films: Comparison of measurement methods,” \nJournal of Applied Physics 99(9), 093909 (2006). \n7 H.T. Nembach, T.J. Silva, J.M. Shaw, M.L. Schneider, M.J. Carey, S. Maat, and J.R. Childress, \n“Perpendicular ferromagnetic resonance measurements of damping and Lande#x0301g - factor in \nsputtered (Co2Mn)1 -xGex thin films,” Phys. Rev. B 84(5), 054424 (2011). \n8 W. Chen, G. de Loubens, J.M.L. Beaujour, A.D. Kent, and J.Z. Sun, “Finite size effects on spin -torque \ndriven ferromagnetic resonance in spin valves with a Co/Ni synthetic free layer,” J. Appl. Phys. 103(7), \n(2008). \n9 D.A. Arena, E. Vescovo, C. -C. Kao, Y . Guan, and W.E. Bailey, “Combined time -resolved x -ray magnetic \ncircular dichroism and ferromagnetic resonance studies of magnetic alloys and multilayers (invited),” \nJournal of Applied Physics 101(9), 09C109 (2007). \n10 G.B.G. Stenning, L.R. Shelford, S.A. Cavill, F. Hoffmann, M. Haertinger, T. Hesjedal, G. Woltersdorf, G.J. \nBowden, S.A. Gregory, C.H. Back, P .A.J. de Groot, and G. van der Laan, “Magnetization dynamics in an exchange -coupled NiFe/CoFe bilayer studied by x -ray detected ferromagnetic resonance,” New J. Phys. \n17(1), 013019 (2015). \n11 D.A. Arena, E. Vescovo, C. -C. Kao, Y . Guan, and W.E. Bailey, “Weakly coupled motion of individual layers \nin ferromagnetic resonance,” Phys. Rev. B 74(6), 064409 (2006). \n12 G. van der Laan, and A.I. Figueroa, “X -ray magnetic circular dichroism —A versatile tool to study \nmagnetism,” Coordination Chemistry Reviews 277–278, 95–129 (2014). \n13 C. Klewe, Q. Li, M. Yang, A.T. N’Diaye, D.M. Burn, T. Hesjedal, A.I. Figueroa, C. Hwang, J. Li, R.J. Hicken, \nP . Shafer, E. Arenholz, G. van der Laan, and Z. Qiu, “Element - and Time -Resolved Measurements of Spin \nDynamics Using X -ray Detected Ferromagnetic Resonance,” Null 33(2), 12 –19 (2020). \n14 M.K. Marcham, L.R. Shelford, S.A. Cavill, P .S. Keatley, W. Yu, P . Shafer, A. Neudert, J.R. Childress, J.A. \nKatine, E. Arenholz, N.D. Telling, G. van der Laan, and R.J. Hicken, “Phase -resolved x -ray ferromagnetic \nresonance measurements of spin pumping in s pin valve structures,” Phys. Rev. B 87(18), 180403 (2013). \n15 P . Warnicke, E. Stavitski, J. -S. Lee, A. Yang, Z. Chen, X. Zuo, S. Zohar, W.E. Bailey, V.G. Harris, and D.A. \nArena, “Direct observation of symmetry -specific precession in a ferrimagnet,” Phys. Rev. B 92(10), \n104402 (2015). \n16 C. Klewe, P . Shafer, J.E. Shoup, C. Kons, Y . Pogoryelov, R. Knut, B.A. Gray, H. -M. Jeon, B.M. Howe, O. \nKaris, Y . Suzuki, E. Arenholz, D.A. Arena, and S. Emori, “Observation of coherently coupled cation spin \ndynamics in an insulating ferrimagnetic oxide,” Applied Physics Letters 122(13), 132401 (2023). \n17 W.E. Bailey, C. Cheng, R. Knut, O. Karis, S. Auffret, S. Zohar, D. Keavney, P . Warnicke, J. -S. Lee, and D.A. \nArena, “Detection of microwave phase variation in nanometre -scale magnetic heterostructures,” Nature \nCommunications 4(1), 2025 (2013). \n18 T. Martin, G. Woltersdorf, C. Stamm, H.A. Dürr, R. Mattheis, C.H. Back, and G. Bayreuther, “Layer \nresolved magnetization dynamics in interlayer exchange coupled Ni81Fe19∕Ru∕Co90Fe10 by time \nresolved x -ray magnetic circular dichroism,” Journal of Applied P hysics 103(7), 07B112 (2008). \n19 G. van der Laan, “Time -resolved X -ray detected ferromagnetic resonance of spin currents,” Journal of \nElectron Spectroscopy and Related Phenomena 220, 137 –146 (2017). \n20 Q. Li, M. Yang, C. Klewe, P . Shafer, A.T. N’Diaye, D. Hou, T.Y . Wang, N. Gao, E. Saitoh, C. Hwang, R.J. \nHicken, J. Li, E. Arenholz, and Z.Q. Qiu, “Coherent ac spin current transmission across an \nantiferromagnetic CoO insulator,” Nat Commun 10(1), 5265 (2019). \n21 J. Li, L.R. Shelford, P . Shafer, A. Tan, J.X. Deng, P .S. Keatley, C. Hwang, E. Arenholz, G. van der Laan, R.J. \nHicken, and Z.Q. Qiu, “Direct Detection of Pure ac Spin Current by X -Ray Pump -Probe Measurements,” \nPhys. Rev. Lett. 117(7), 076602 (2016). \n22 M. Dąbrowski, T. Nakano, D.M. Burn, A. Frisk, D.G. Newman, C. Klewe, Q. Li, M. Yang, P . Shafer, E. \nArenholz, T. Hesjedal, G. van der Laan, Z.Q. Qiu, and R.J. Hicken, “Coherent Transfer of Spin Angular \nMomentum by Evanescent Spin Waves within Antiferromagn etic NiO,” Phys. Rev. Lett. 124(21), 217201 \n(2020). \n23 Y . Pogoryelov, M. Pereiro, S. Jana, A. Kumar, S. Akansel, M. Ranjbar, D. Thonig, D. Primetzhofer, P . \nSvedlindh, J. Åkerman, O. Eriksson, O. Karis, and D.A. Arena, “Nonreciprocal spin pumping damping in \nasymmetric magnetic trilayers,” Phys. Rev. B 101(5), 054401 (2020). \n24 T. Schaffers, T. Feggeler, S. Pile, R. Meckenstock, M. Buchner, D. Spoddig, V. Ney, M. Farle, H. Wende, S. \nWintz, M. Weigand, H. Ohldag, K. Ollefs, and A. Ney, “Extracting the Dynamic Magnetic Contrast in Time -\nResolved X -Ray Transmission Microscopy,” Nano materials 9(7), 940 (2019). \n25 M.G. Silly, T. Ferté, M.A. Tordeux, D. Pierucci, N. Beaulieu, C. Chauvet, F. Pressacco, F. Sirotti, H. \nPopescu, V. Lopez -Flores, M. Tortarolo, M. Sacchi, N. Jaouen, P . Hollander, J.P . Ricaud, N. Bergeard, C. \nBoeglin, B. Tudu, R. Delaunay, J. Luning, G. Ma linowski, M. Hehn, C. Baumier, F. Fortuna, D. Krizmancic, L. \nStebel, R. Sergo, and G. Cautero, “Pump−probe experiments at the TEMPO beamline using the low -α \noperation mode of Synchrotron SOLEIL,” J Synchrotron Rad 24(4), 886 –897 (2017). 26 G. Materlik, T. Rayment, and D.I. Stuart, “Diamond Light Source: status and perspectives,” \nPhilosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences \n373(2036), 20130161 (2015). \n27 S. Pile, S. Stienen, K. Lenz, R. Narkowicz, S. Wintz, J. Förster, S. Mayr, M. Buchner, M. Weigand, V. Ney, J. \nLindner, and A. Ney, “Nonstationary spin waves in a single rectangular permalloy microstrip under \nuniform magnetic excitation,” Phys. Rev. B 105(9), 094415 (2022). \n28 S. Bonetti, R. Kukreja, Z. Chen, D. Spoddig, K. Ollefs, C. Schöppner, R. Meckenstock, A. Ney, J. Pinto, R. \nHouanche, J. Frisch, J. Stöhr, H.A. Dürr, and H. Ohldag, “Microwave soft x -ray microscopy for nanoscale \nmagnetization dynamics in the 5 –10 GHz frequ ency range,” Review of Scientific Instruments 86(9), \n093703 (2015). \n29 S. Mizukami, F. Wu, A. Sakuma, J. Walowski, D. Watanabe, T. Kubota, X. Zhang, H. Naganuma, M. \nOogane, Y . Ando, and T. Miyazaki, “Long -Lived Ultrafast Spin Precession in Manganese Alloys Films with a \nLarge Perpendicular Magnetic Anisotropy,” Phys. Rev. Lett . 106(11), 117201 (2011). \n30 S. Mizukami, S. Iihama, N. Inami, T. Hiratsuka, G. Kim, H. Naganuma, M. Oogane, and Y . Ando, “Fast \nmagnetization precession observed in L10 -FePt epitaxial thin film,” Applied Physics Letters 98(5), \n052501 -052501 –3 (2011). \n31 J.M. Shaw, H.T. Nembach, and T.J. Silva, “Determination of spin pumping as a source of linewidth in \nsputtered Co${}_{90}$Fe${}_{10}$/Pd multilayers by use of broadband ferromagnetic resonance \nspectroscopy,” Phys. Rev. B 85(5), 054412 (2012). \n32 M. Jaris, W. Yang, C. Berk, and H. Schmidt, “Towards ultraefficient nanoscale straintronic microwave \ndevices,” Phys. Rev. B 101(21), 214421 (2020). \n33 A. Barman, S. Wang, O. Hellwig, A. Berger, E.E. Fullerton, and H. Schmidt, “Ultrafast magnetization \ndynamics in high perpendicular anisotropy [Co∕Pt][sub n] multilayers,” Journal of Applied Physics 101, \n09D102 (2007). \n34 S. Wei, and M.Y . Chou, “Phonon dispersions of silicon and germanium from first -principles \ncalculations,” Phys. Rev. B 50(4), 2221 –2226 (1994). \n35 T. Popmintchev, M. -C. Chen, P . Arpin, M.M. Murnane, and H.C. Kapteyn, “The attosecond nonlinear \noptics of bright coherent X -ray generation,” Nature Photon 4(12), 822 –832 (2010). \n36 C. La -O-Vorakiat, M. Siemens, M.M. Murnane, H.C. Kapteyn, S. Mathias, M. Aeschlimann, P . Grychtol, \nR. Adam, C.M. Schneider, J.M. Shaw, H. Nembach, and T.J. Silva, “Ultrafast Demagnetization Dynamics at \nthe $M$ Edges of Magnetic Elements Observed Using a T abletop High -Harmonic Soft X -Ray Source,” \nPhys. Rev. Lett. 103(25), 257402 (2009). \n37 S. Mathias, C. La -O-Vorakiat, P . Grychtol, P . Granitzka, E. Turgut, J.M. Shaw, R. Adam, H.T. Nembach, \nM.E. Siemens, S. Eich, C.M. Schneider, T.J. Silva, M. Aeschlimann, M.M. Murnane, and H.C. Kapteyn, \n“Probing the timescale of the exchange interaction in a ferromagnetic alloy,” PNAS 109(13), 4792 –4797 \n(2012). \n38 D. Rudolf, C. La -O-Vorakiat, M. Battiato, R. Adam, J.M. Shaw, E. Turgut, P . Maldonado, S. Mathias, P . \nGrychtol, H.T. Nembach, T.J. Silva, M. Aeschlimann, H.C. Kapteyn, M.M. Murnane, C.M. Schneider, and \nP .M. Oppeneer, “Ultrafast magnetization enhancement i n metallic multilayers driven by superdiffusive \nspin current,” Nat Commun 3, 1037 (2012). \n39 E. Turgut, D. Zusin, D. Legut, K. Carva, R. Knut, J.M. Shaw, C. Chen, Z. Tao, H.T. Nembach, T.J. Silva, S. \nMathias, M. Aeschlimann, P .M. Oppeneer, H.C. Kapteyn, M.M. Murnane, and P . Grychtol, “Stoner versus \nHeisenberg: Ultrafast exchange reduction and mag non generation during laser -induced \ndemagnetization,” Phys. Rev. B 94(22), 220408 (2016). \n40 M. Hofherr, S. Häuser, J.K. Dewhurst, P . Tengdin, S. Sakshath, H.T. Nembach, S.T. Weber, J.M. Shaw, T.J. \nSilva, H.C. Kapteyn, M. Cinchetti, B. Rethfeld, M.M. Murnane, D. Steil, B. Stadtmüller, S. Sharma, M. \nAeschlimann, and S. Mathias, “Ultrafast opticall y induced spin transfer in ferromagnetic alloys,” Science \nAdvances 6(3), eaay8717 (2020). 41 P . Tengdin, C. Gentry, A. Blonsky, D. Zusin, M. Gerrity, L. Hellbrück, M. Hofherr, J. Shaw, Y . Kvashnin, E.K. \nDelczeg -Czirjak, M. Arora, H. Nembach, T.J. Silva, S. Mathias, M. Aeschlimann, H.C. Kapteyn, D. Thonig, K. \nKoumpouras, O. Eriksson, and M.M. Murn ane, “Direct light –induced spin transfer between different \nelements in a spintronic Heusler material via femtosecond laser excitation,” Science Advances 6(3), \neaaz1100 (2020). \n42 S. Jana, J.A. Terschlüsen, R. Stefanuik, S. Plogmaker, S. Troisi, R.S. Malik, M. Svanqvist, R. Knut, J. \nSöderström, and O. Karis, “A setup for element specific magnetization dynamics using the transverse \nmagneto -optic Kerr effect in the energy range of 30 -72 eV,” Review of Scientific Instruments 88(3), \n033113 (2017). \n43 H. Liu, R. Knut, S. Saha, R.S. Malik, K. Jatkar, R. Stefanuik, J. Söderström, J.E. Shoup, D. Khadka, T.R. \nThapaliya, S.X. Huang, A. Gupta, O. Karis, D. Karaiskaj, and D.A. Arena, “Optical and extreme UV studies \nof spin dynamics in metallic and insulating ferrimagnets,” Journal of Applied Physics 130(24), 240901 \n(2021). \n44 I. Vaskivskyi, R.S. Malik, L. Salemi, D. Turenne, R. Knut, J. Brock, R. Stefanuik, J. Söderström, K. Carva, \nE.E. Fullerton, P .M. Oppeneer, O. Karis, and H.A. Dürr, “Element -Specific Magnetization Dynamics in Co –\nPt Alloys Induced by Strong Optical Excitati on,” J. Phys. Chem. C 125(21), 11714 –11721 (2021). \n45 J.L. Ellis, K.M. Dorney, D.D. Hickstein, N.J. Brooks, C. Gentry, C. Hernández -García, D. Zusin, J.M. Shaw, \nQ.L. Nguyen, C.A. Mancuso, G.S.M. Jansen, S. Witte, H.C. Kapteyn, and M.M. Murnane, “High harmonics \nwith spatially varying ellipticity,” Optica, OPT ICA 5(4), 479 –485 (2018). \n46 D.D. Hickstein, F.J. Dollar, P . Grychtol, J.L. Ellis, R. Knut, C. Hernández -García, D. Zusin, C. Gentry, J.M. \nShaw, T. Fan, K.M. Dorney, A. Becker, A. Jaroń -Becker, H.C. Kapteyn, M.M. Murnane, and C.G. Durfee, \n“Non -collinear generation of angularly isolat ed circularly polarized high harmonics,” Nature Photonics \n9(11), 743 (2015). \n47 O. Kfir, P . Grychtol, E. Turgut, R. Knut, D. Zusin, D. Popmintchev, T. Popmintchev, H. Nembach, J.M. \nShaw, A. Fleischer, H. Kapteyn, M. Murnane, and O. Cohen, “Generation of bright phase -matched \ncircularly -polarized extreme ultraviolet high harmonics,” Na t Photon 9(2), 99 –105 (2015). \n48 K. Yao, F. Willems, C. von Korff Schmising, I. Radu, C. Strüber, D. Schick, D. Engel, A. Tsukamoto, J.K. \nDewhurst, S. Sharma, and S. Eisebitt, “Distinct spectral response in $M$ -edge magnetic circular \ndichroism,” Phys. Rev. B 102(10), 100405 (2020). \n49 B. Vodungbo, A.B. Sardinha, J. Gautier, G. Lambert, C. Valentin, M. Lozano, G. Iaquaniello, F. Delmotte, \nS. Sebban, J. Lüning, and P . Zeitoun, “Polarization control of high order harmonics in the EUV photon \nenergy range,” Opt. Express, OE 19(5), 4346 –4356 (2011). \n50 B.L. Henke, E.M. Gullikson, and J.C. Davis, “X -Ray Interactions: Photoabsorption, Scattering, \nTransmission, and Reflection at E = 50 -30,000 eV, Z = 1 -92,” Atomic Data and Nuclear Data Tables 54(2), \n181–342 (1993). \n51 J. Miao, T. Ishikawa, I.K. Robinson, and M.M. Murnane, “Beyond crystallography: Diffractive imaging \nusing coherent x -ray light sources,” Science 348(6234), 530 –535 (2015). \n52 J. Miao, P . Charalambous, J. Kirz, and D. Sayre, “Extending the methodology of X -ray crystallography to \nallow imaging of micrometre -sized non -crystalline specimens,” Nature 400(6742), 342 –344 (1999). \n53 N. Bukin, C. McKeever, E. Burgos -Parra, P .S. Keatley, R.J. Hicken, F.Y . Ogrin, G. Beutier, M. Dupraz, H. \nPopescu, N. Jaouen, F. Yakhou -Harris, S.A. Cavill, and G. van der Laan, “Time -resolved imaging of \nmagnetic vortex dynamics using holography with exten ded reference autocorrelation by linear \ndifferential operator,” Sci Rep 6(1), 36307 (2016). \n54 D.M. Burn, S.L. Zhang, G.Q. Yu, Y . Guang, H.J. Chen, X.P . Qiu, G. van der Laan, and T. Hesjedal, “Depth -\nResolved Magnetization Dynamics Revealed by X -Ray Reflectometry Ferromagnetic Resonance,” Phys. \nRev. Lett. 125(13), 137201 (2020). \n55 M. Tanksalvala, C.L. Porter, Y . Esashi, B. Wang, N.W. Jenkins, Z. Zhang, G.P . Miley, J.L. Knobloch, B. \nMcBennett, N. Horiguchi, S. Yazdi, J. Zhou, M.N. Jacobs, C.S. Bevis, R.M. Karl, P . Johnsen, D. Ren, L. Waller, D.E. Adams, S.L. Cousin, C. -T. Liao, J. Miao, M. Gerrity, H.C. Kapteyn, and M.M. Murnane, \n“Nondestructive, high -resolution, chemically specific 3D nanostructure characterization using phase -\nsensitive EUV imaging reflectometry,” Science Advances 7(5), eabd9667 (2021). \n56 M.A.W. Schoen, D. Thonig, M.L. Schneider, T.J. Silva, H.T. Nembach, O. Eriksson, O. Karis, and J.M. Shaw, \n“Ultra -low magnetic damping of a metallic ferromagnet,” Nat Phys 12(9), 839 –842 (2016). \n57 E.R.J. Edwards, H.T. Nembach, and J.M. Shaw, “${ \\mathrm{Co}}_{25}{ \\mathrm{Fe}}_{75}$ Thin Films \nwith Ultralow Total Damping of Ferromagnetic Resonance,” Phys. Rev. Applied 11(5), 054036 (2019). \n58 M.A.W. Schoen, J. Lucassen, H.T. Nembach, B. Koopmans, T.J. Silva, C.H. Back, and J.M. Shaw, \n“Magnetic properties in ultrathin $3d$ transition -metal binary alloys. II. Experimental verification of \nquantitative theories of damping and spin pumping,” Phys. Rev. B 95(13), 134411 (2017). \n59 P .C. Johnsen, S.A. Ryan, C. Gentry, A. Grafov, H. Kapteyn, and M. Murnane, “A beamline for ultrafast \nextreme ultraviolet magneto -optical spectroscopy in reflection near the shot noise limit,” Review of \nScientific Instruments 94(3), 033001 (2023). \n60 Y . Esashi, N.W. Jenkins, Y . Shao, J.M. Shaw, S. Park, M.M. Murnane, H.C. Kapteyn, and M. Tanksalvala, \n“Tabletop extreme ultraviolet reflectometer for quantitative nanoscale reflectometry, scatterometry, and \nimaging,” Review of Scientific Instruments 94(12), 123705 (2023). \n61 K. Yao, F. Willems, C. von Korff Schmising, C. Strüber, P . Hessing, B. Pfau, D. Schick, D. Engel, K. \nGerlinger, M. Schneider, and S. Eisebitt, “A tabletop setup for ultrafast helicity -dependent and element -\nspecific absorption spectroscopy and scattering i n the extreme ultraviolet spectral range,” Review of \nScientific Instruments 91(9), 093001 (2020). \n62 J. Zak, E.R. Moog, C. Liu, and S.D. Bader, “Magneto -optics of multilayers with arbitrary magnetization \ndirections,” Phys. Rev. B 43(8), 6423 –6429 (1991). \n63 T.M. Fortier, A. Rolland, F. Quinlan, F.N. Baynes, A.J. Metcalf, A. Hati, A.D. Ludlow, N. Hinkley, M. \nShimizu, T. Ishibashi, J.C. Campbell, and S.A. Diddams, “Optically referenced broadband electronic \nsynthesizer with 15 digits of resolution,” Laser & Pho tonics Reviews 10(5), 780 –790 (2016). \n64 A. Bartels, S.A. Diddams, C.W. Oates, G. Wilpers, J.C. Bergquist, W.H. Oskay, and L. Hollberg, \n“Femtosecond -laser -based synthesis of ultrastable microwave signals from optical frequency references,” \nOpt. Lett., OL 30(6), 667 –669 (2005). \n65 A. Bartels, S.A. Diddams, T.M. Ramond, and L. Hollberg, “Mode -locked laser pulse trains with \nsubfemtosecond timing jitter synchronized to an optical reference oscillator,” Opt. Lett., OL 28(8), 663 –\n665 (2003). \n66 M. Holler, M. Guizar -Sicairos, E.H.R. Tsai, R. Dinapoli, E. Müller, O. Bunk, J. Raabe, and G. Aeppli, “High -\nresolution non -destructive three -dimensional imaging of integrated circuits,” Nature 543(7645), 402 –406 \n(2017). \n67 A. Rana, C. -T. Liao, E. Iacocca, J. Zou, M. Pham, E. -E.C. Subramanian, Y .H. Lo, S.A. Ryan, X. Lu, C.S. Bevis, \nR.M. Karl Jr, A.J. Glaid, Y . -S. Yu, P . Mahale, D.A. Shapiro, S. Yazdi, T.E. Mallouk, S.J. Osher, H.C. Kapteyn, \nV.H. Crespi, J.V. Badding, Y . Tser kovnyak, M.M. Murnane, and J. Miao, “Direct observation of 3D \ntopological spin textures and their interactions using soft x -ray vector ptychography,” (2021). \n " }, { "title": "2101.05831v1.Anomalous_Hall_effect_in_weak_itinerant_ferrimagnet_FeCr__2_Te__4_.pdf", "content": "arXiv:2101.05831v1 [cond-mat.str-el] 14 Jan 2021Anomalous Hall effect in weak-itinerant ferrimagnet FeCr 2Te4\nYu Liu,1,∗Hengxin Tan,2Zhixiang Hu,1,3Binghai Yan,2and C. Petrovic1,3\n1Department of Condensed Matter Physics and Materials Scien ce,\nBrookhaven National Laboratory, Upton, New York 11973, USA\n2Department of Condensed Matter Physics, Weizmann Institut e of Science, Rehovot 7610001, Israel\n3Department of Materials Science and Chemical Engineering,\nStony Brook University, Stony Brook, NY 11790, USA\n(Dated: January 18, 2021)\nWe carried out a comprehensive study of electronic transpor t, thermal and thermodynamic prop-\nerties inFeCr 2Te4single crystals. Itexhibits bad-metallic behavior andano malous Hall effect(AHE)\nbelow a weak-itinerant paramagentic-to-ferrimagnetic tr ansition Tc∼123 K. The linear scaling be-\ntween the anomalous Hall resistivity ρxyand the longitudinal resistivity ρxximplies that the AHE\nin FeCr 2Te4is most likely dominated by extrinsic skew-scattering mech anism rather than intrinsic\nKL or extrinsic side-jump mechanism, which is supported by o ur Berry phase calculations.\nINTRODUCTION\nThe anomalous Hall effect (AHE) in metals is linked\nto an asymmetry in carrier paths and the effects of spin-\norbit interaction. This is typically observed in ferromag-\nnets since an electric current induces a transversevoltage\ndrop in zero magnetic field which is proportional to mag-\nnetization[1,2]. Spin-orbitcouplinginthe ferromagnetic\nbands leads to anomalous carrier velocities and intrinsic\nAHE [3]. The intrinsic Kaplus-Luttinger (KL) mecha-\nnism can be reinterpreted as a manifestation of Berry-\nphase effects on occupied electronic Bloch states [4, 5].\nThe extrinsic mechanisms involving skew-scattering and\nside-jump mechanisms can also give rise to the AHE and\nare induced by asymmetric scattering of conduction elec-\ntrons [6, 7]. In recent years it has been shown that the\nAHEvelocitiesarisefromthetopologicalBerrycurvature\nwhich generate an effective magnetic field in momentum\nspace in varieties of Dirac materials with noncollinear\nspin configuration [8–12].\nFeCr2Ch4(Ch = O, S, Se, Te) materials show rich cor-\nrelatedelectronphysics. FeCr 2O4spinelshowsacomplex\nmagnetic phase diagram with a ferrimagnetic (FIM) and\nmultiferroic order below 80 K, a strong spin-lattice cou-\npling and orbital order due to the Jahn-Teller distortion\n[13–17]. FeCr 2S4isa multiferroicferrimagnetbelow Tc=\n165 K with large changes of resistivity in magnetic field\n[18–22]. FeCr 2Se4orders antiferromagnetically with TN\n= 218 K in an insulating state despite with larger ligand\nchalcogenatom[23–25]. FeCr 2S4andFeCr 2Se4havesim-\nilar electronic structure with nearly trivalent Cr3+and\ndivalent Fe2+states, and there is a strong hybridization\nbetween Fe 3 d- and Ch p-states [26]. FeCr 2Te4shows no\nsemiconducting gap and a FIM order below Tc= 123 K\n[27, 28].\nIn this work, we performed a comprehensive study of\nelectronic and thermal transport properties in FeCr 2Te4\nsinglecrystals. TheAHEobservedbelow Tcisdominated\nby the skew-scattering mechanism, i.e., by the Bloch\nstate transport lifetime arising from electron scatteringby impurities or defects in the presence of spin-orbit ef-\nfects, and is smaller than the intrinsic AHE revealed by\ndensity functional calculations.\nEXPERIMENTAL AND COMPUTATIONAL\nDETAILS\nSingle crystals growth and crystal structure details\nare described in ref.[28]. Electrical and thermal trans-\nport were measured in quantum design PPMS-9. The\nlongitudinal and Hall resistivity were measured using\na standard four-probe method. In order to effectively\neliminate the longitudinal resistivity contribution due to\nvoltage probe misalignment, the Hall resistivity was ob-\ntained by the difference of transverse resistance mea-\nsured at positive and negative fields, i.e., ρxy(µ0H) =\n[ρ(+µ0H)−ρ(−µ0H)]/2. Isothermal magnetization was\nmeasured in quantum design MPMS-XL5.\nWe performed density functional theory (DFT) cal-\nculations with the Perdew-Burke-Ernzerhof (PBE) [29]\nexchange-correlation functional that is implemented in\nthe Vienna ab initio simulation package(VASP) [30]. We\nadoptedtheexperimentalcrystalstructurewiththe ferri-\nmagnetism (parallelto the lattice vector c) [28]. The cut-\noff energy for the plane wavebasis is 300eV. A k-mesh of\n10×10×10wasused in the Brillouin zone sampling. The\nspin-orbit coupling was included. The intrinsic anoma-\nlousHallconductivity(AHC) andSeebeckcoefficientwas\ncalculated in a tight-binding scheme based on the maxi-\nmally localized Wannier functions [31].\nRESULTS AND DISCUSSIONS\nFigure 1(a) shows the temperature-dependent heat ca-\npacityCp(T) for FeCr 2Te4. A clear anomaly around 123\nK corresponds well to the paramagnetic (PM)-FIM tran-\nsition. The high temperature Cp(T) approaches the Du-\nlong Petit value of 3 NR≈172 J mol−1K−1, whereR2\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s52/s48/s56/s48/s49/s50/s48/s49/s54/s48\n/s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s50/s52/s54/s56/s49/s48/s48 /s49/s50/s48 /s50/s52/s48 /s51/s54/s48/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s50/s52/s54\n/s48 /s53/s48 /s49/s48/s48/s52/s46/s48/s52/s46/s53/s53/s46/s48/s67\n/s112/s32/s40/s74/s47/s109/s111/s108/s45/s75/s41\n/s84/s32/s40/s75/s41/s40/s97/s41\n/s51/s78/s82/s49/s50/s51/s32/s75/s83/s32/s40 /s86/s47/s75/s41\n/s84/s32/s40/s75/s41/s40/s98/s41\n/s67\n/s112/s47/s84/s32/s40/s74/s47/s109/s111/s108/s45/s75/s50\n/s41\n/s84/s50\n/s32/s40/s75/s50\n/s41\n/s40/s99/s41/s120/s120/s32/s40/s49/s48/s45/s52\n/s32/s99/s109/s41\n/s84/s32/s40/s75/s41/s120/s120/s40/s49/s48/s45/s52\n/s99/s109/s41\n/s84/s32/s40/s75/s41\nFIG. 1. (Color online) (a) Temperature-dependent heat ca-\npacityCp(T) for FeCr 2Te4. Inset shows the low temperature\nCp(T)/TvsT2curve fitted by Cp(T)/T=γ+βT2. (b) See-\nbeck coefficient S(T) and (c) in-plane resistivity ρxx(T) for\nFeCr2Te4single crystal. Inset in (c) shows data below 100 K\nfitted by ρ(T) =ρ0+aT3/2+bT2(solid line) in comparison\nwithρ(T) =ρ0+cT2(dashed line).\n= 8.314 J mol−1K−1is the molar gas constant. The\nlow temperature data from 2 to 18 K are featureless and\ncould be fitted by using Cp(T)/T=γ+βT2, where the\nfirst term is the Sommerfeld electronic specific heat co-\nefficient and the second term is low-temperature limit of\nlattice heat capacity[inset in Fig. 1(a)]. The fitting gives\nγ= 61(2) mJ mol−1K−2andβ= 1.7(1) mJ mol−1K−4.\nThe Debye temperature Θ D= 199(1) K can be calcu-\nlated by using Θ D= (12π4NR/5β)1/3, whereN= 7 is\nthe number of atoms per formula unit.\nThe Seebeck coefficient S(T) of FeCr 2Te4is positive in\nthe whole temperature range, indicating dominant hole-\ntype carriers [Fig. 1(b)]. The S(T) changes slope around\nTcand gradually decreases with decreasing temperature.\nAs we know, the S(T) depends sensitively on the Fermi\nsurface. The slope change of S(T) reflects the possible\nreconstruction of Fermi surface passing through the PM-\nFIM transition. At low temperature, the diffusive See-\nbeck response of Fermi liquid dominates and is expected\nto be linear in T. In a metal with dominant single-band\ntransport, the Seebeck coefficient could be described by\nthe Mott relationship,\nS=π2\n3k2\nBT\neN(εF)\nn, (1)\nwhereN(εF) is the density of states (DOS), εFis the\nFermi energy, nis carrier concentration, kBis the Boltz-\nman constant and eis the absolute value of electronic\ncharge [32]. The derived dS/dTbelow 26 K is ∼0.074(1)\nµV K−2. TheS(T) curve is consistent with our calcu-/s48 /s50 /s52 /s54 /s56/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52\n/s48 /s49 /s50 /s51 /s52 /s53/s48/s49/s50/s51/s52\n/s114\n/s120/s121/s32/s40 /s109/s87 /s32/s99/s109/s41\n/s109\n/s48/s72/s32/s40/s84/s41/s32/s56/s48/s32/s75\n/s32/s49/s48/s48/s32/s75\n/s32/s49/s50/s48/s32/s75/s40/s98/s41\n/s109\n/s48/s72/s32/s47/s47/s32/s99\n/s32/s50/s48/s32/s75\n/s32/s52/s48/s32/s75\n/s32/s54/s48/s32/s75/s77/s32/s40 /s109\n/s66/s47/s102/s46/s117/s46/s41\n/s109\n/s48/s72/s40/s84/s41/s32/s50/s48/s32/s75\n/s32/s52/s48/s32/s75\n/s32/s54/s48/s32/s75\n/s32/s56/s48/s32/s75\n/s32/s49/s48/s48/s32/s75\n/s32/s49/s50/s48/s32/s75/s40/s97/s41\n/s109\n/s48/s72/s32/s47/s47/s32/s99\n/s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48 /s49/s50/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53\n/s84/s32/s40/s75/s41/s82\n/s48/s32/s40/s49/s48/s45/s50\n/s32/s99/s109/s51\n/s47/s67/s41/s40/s99/s41\n/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53\n/s82\n/s115/s32/s40/s99/s109/s51\n/s47/s67/s41\nFIG. 2. (Color online) Out-of-plane field dependence of (a) d c\nmagnetization M(µ0H) and (b) Hall resistivity ρxy(µ0H) for\nFeCr2Te4at indicated temperatures. (c) Temperature depen-\ndence of ordinary Hall coefficient R0(left axis) and anoma-\nlous Hall coefficient Rs(right axis) fitted from the ρxyvsµ0H\ncurves using ρxy=R0µ0H+RsM.\nlations based on Boltzmann equations and DFT band\nstructure [see below in Fig. 4(b)]. The electronic specific\nheat is:\nCe=π2\n3k2\nBTN(εF). (2)\nFrom Eq. (1), thermopower probes the specific heat per\nelectron: S=Ce/ne. The units are V K−1forS, J K−1\nm−3forCe, and m−3forn, respectively. It is common to\nexpressγ=Ce/TinJK−2mol−1units. Inordertofocus\non theS/Ceratio, we define a dimensionless quantity\nq=S\nTNAe\nγ, (3)\nwhereNAis the Avogadro number. This gives the num-\nber of carriers per formula unit (proportional to 1 /n)\n[33]. The obtained q= 0.10(1) indicates about 0.1 hole\nper formula unit within the Boltzmann framework [33].\nFigure 1(c) shows the temperature-dependent in-plane\nresistivity ρxx(T) of FeCr 2Te4, indicating a metallic be-\nhavior with a relatively low residual resistivity ratio\n[RRR = ρ(300K)/ ρ(2K) =1.7]. Aclearkinkis observed\natTc, correspondingwell to the PM-FIM transition. The\nrenormalized spin fluctuation theory suggests that the\nelectrical resistivity shows a T2dependence for itinerant\nferromagnetic system [34]. In FeCr 2Te4, the low temper-\nature resistivity fitting gives a better result by adding an3\n/s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48 /s49/s50/s48/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53\n/s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48 /s49/s50/s48/s48/s50/s52/s54/s56/s49/s48/s50/s48 /s52/s48 /s54/s48 /s56/s48 /s49/s48/s48 /s49/s50/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48\n/s45/s51/s46/s52/s48 /s45/s51/s46/s51/s53 /s45/s51/s46/s51/s48/s45/s53/s46/s53/s53/s45/s53/s46/s53/s48/s45/s53/s46/s52/s53/s45/s53/s46/s52/s48\n/s65 /s120/s121/s32/s40\n/s99/s109/s41/s45/s49\n/s84/s32/s40/s75/s41/s40/s98/s41\n/s40/s100/s41/s40/s97/s41\n/s83\n/s72/s32/s40/s49/s48/s45/s50\n/s32/s86/s45/s49\n/s41\n/s84/s32/s40/s75/s41/s110/s32/s40/s49/s48/s50/s49\n/s32/s99/s109/s45/s51\n/s41\n/s84/s32/s40/s75/s41/s108/s111/s103/s32/s65 /s120/s121\n/s32/s40 /s99/s109/s41\n/s108/s111/s103/s32\n/s120/s120/s32/s40 /s32/s99/s109/s41/s32/s61/s32/s49/s46/s49/s40/s50/s41/s40/s99/s41\nFIG. 3. (Color online) Temperature dependence of the carrie r\nconcentration (a) and the anomalous Hall conductivity σA\nxy=\nρA\nxy/(ρ2\nxx+ρ2\nxy) (b). Scaling behavior of the anomalous Hall\nresistivity (c) and the coefficient SH=µ0Rs/ρ2\nxx(d).\nadditional T3/2term that describes the contribution of\nspin fluctuation scattering [35].\nρ(T) =ρ0+aT3\n2+bT2, (4)\nwhereρ0is the residual resistivity, aandbare constants.\nThe fitting yields ρ0= 366(1) µΩ cm,a= 1.00(3) ×10−1\nµΩ cm K−1, andb= 2.8(3)×10−3µΩ cm K−2, indicating\ntheT3/2term predominates. This means the interaction\nbetween conduction electrons and localized spins could\nnot be simply treated asa small perturbation to a system\nof free electrons, i.e., strong electron correlation should\nbe considered in FeCr 2Te4.\nFigure 2(a) shows the isothermal magnetization mea-\nsured at varioustemperatures below Tc. All the M(µ0H)\ncurves rapidly increase in low field and change slowly in\nhigh field. Field dependence of Hall resistivity ρxy(µ0H)\nfor FeCr 2Te4at the corresponding temperatures are de-\npicted in Fig. 2(b). All the ρxy(µ0H) curves jump in low\nfield and then become linear-in-field in high field, indi-\ncating an AHE in FeCr 2Te4crystal. In general, the Hall\nresistivity ρxyin ferromagnets is made up of two parts,\nρxy=ρO\nxy+ρA\nxy=R0µ0H+RsM, (5)\nwhereρO\nxyandρA\nxyare the ordinary and anomalous Hall\nresistivity, respectively [36–39]. R0is the ordinary Hall\ncoefficientfromwhichapparentcarrierconcentrationand\ntype can be determined ( R0= 1/nq).Rsis the anoma-\nlous Hall coefficient. With a linear fit of ρxy(µ0H) in\nFIG. 4. (Color online) (a) Crystal structure, Brillouin zon e\n(BZ),andelectronic structureofFeCr 2Te4. Theredvectorsin\nthe crystal structure represent the directions of the magne tic\nmoments on Fe and Cr. The high symmetric k-paths in the\nBZ are shown. The x, y, and z directions of the Cartesian co-\nordinate are along the lattice vectors a, b, and c, respectiv ely.\nThe Fermi energy is set tozero. Calculated (b)Seebeck coeffi-\ncientSand (c) anomalous Hall conductivity σxyof FeCr 2Te4.\nThe calculated Sin low temperature shows good agreement\nwith the experiment. The σxyat the Fermi level (zero) is ∼\n127 (Ω cm)−1, much larger than the measured value of 22 (Ω\ncm)−1.\nhigh field, the slope and intercept corresponds to R0and\nρA\nxy, respectively. Rscan be obtained from ρA\nxy=RsMs\nwithMstaken from linear fit of M(µ0H) curves in high\nfield. The temperature dependence of derived R0andRs\nis plotted in Fig. 2(c). The value of R0is positive, in line\nwith the positive S(T), confirming the hole-type carries.\nThe derived Rsgradually decreases with decreasing tem-\nperature. Its magnitude is about two orders larger than\nthat ofR0.\nThe derived carrier concentration nis shown in Fig.\n3(a). The n∼0.5×1021cm−3at 20 K corresponds to ∼\n0.04 holes per formula unit, comparable to the value es-\ntimated from q. Taken into account a weak temperature-\ndependent ρ(T) [Fig. 1(c)], the estimated n∼1.11×1021\ncm−3from 484 µΩ cm near 100 K points to a mean free\npathλ∼0.44 nm. This is comparable to the lattice pa-\nrameters and is close to the Mott-Ioffe-Regel limit [40].\nThe AHC σA\nxy(≈ρA\nxy/ρ2\nxx) is plotted in Fig. 3(b). Theo-\nretically, intrinsic contribution of σA\nxy,inis of the order of\ne2/(hd), where eis the electronic charge, his the Plank\nconstant, and dis the lattice parameter [41]. Taking\nd≈V1/3∼4.3˚A,σA\nxy,inis estimated ∼900 (Ω cm)−1,4\nmuch larger than the obtained values in Fig. 3(b). Ex-\ntrinsic side-jump contribution of σA\nxy,sjis usually of the\norder of e2/(hd)(εSO/EF), where εSOandEFis spin-\norbital interaction energy and Fermi energy, respectively\n[42]. The value of εSO/EFis generally less than 10−2for\nmetallic ferromagnets. As we can see, the σA\nxyis about\n22 (Ω cm)−1at 20 K and exhibits a moderate tempera-\nture dependence. This value is much smaller than σA\nxy,in\n∼900 (Ω cm)−1, which precludes the possibility of in-\ntrinsic KL mechanism. Based on the band structure, as\nshown in Fig. 4, we obtained the intrinsic AHC as127(Ω\ncm)−1,whichismuchlargerthanthemeasuredvaluetoo.\nThe extrinsic side-jump mechanism, where the potential\nfield induced by impurities contributes to the anomalous\ngroup velocity, follows a scaling behavior of ρA\nxy=βρ2\nxx,\nthe same with intrinsic KL mechanism. The scaling be-\nhaviorof ρA\nxyvsρxxgivesα∼1.1(2) by using ρA\nxy=βρα\nxx\n[Fig. 3(c)], which also precludes the possibility of side-\njump and KL mechanism with α= 2. It points to that\nthe skew-scattering possibly dominates, which describes\nasymmetric scattering induced by impurities or defects\nand contributes to AHE with α= 1. Furthermore, the\nscaling coefficient SH=µ0Rs/ρ2\nxx=σA\nxy/Ms[Fig. 3(d)]\nis weaklytemperature-dependent and iscomparablewith\nthose in traditional itinerant ferromagnets, such as Fe\nand Ni (SH∼0.01−0.2V−1) [43, 44]. It is proposedthat\nthe FIM in FeCr 2Te4is itinerant ferromagnetism among\nantiferromagnetically coupled Cr-Fe-Cr trimers [28]. In\na noncomplanar spin trimer structures the topologically\nnontrivial Berry phase is induced by spin chirality rather\nthan spin-orbit effect, resulting in chirality-induced in-\ntrinsic AHE [45–48]. Our result excludes such scenario\nin Cr-Fe-Cr trimers in FeCr 2Te4[28].\nCONCLUSIONS\nIn summary, we studied the electronic transport prop-\nerties and AHE in FeCr 2Te4single crystal. The AHE be-\nlowTc= 123 K is dominated by extrinsic skew-scattering\nmechanism rather than the intrinsic KL or extrinsic side-\njump mechanism, which is confirmed by our DFT calcu-\nlations. The spin structure of Cr-Fe-Cr trimers proposed\nfor FeCr 2Te4is of interest to check by neutron scattering\nexperiments on powder and single crystals in the future.\nACKNOWLEDGEMENTS\nWork at Brookhaven National Laboratory (BNL) is\nsupportedbytheOfficeofBasicEnergySciences, Materi-\nals Sciences and Engineering Division, U.S. Department\nof Energy (DOE) under Contract No. DE-SC0012704.\nB.Y. acknowledges the financial support by the Willner\nFamily Leadership Institute for the Weizmann Institute\nof Science, the Benoziyo Endowment Fund for the Ad-vancement of Science, Ruth and Herman Albert Schol-\nars Program for New Scientists, the European Research\nCouncil (ERC Consolidator Grant No. 815869, “Nonlin-\nearTopo”).\n∗Present address: Los Alamos National Laboratory,\nMS K764, Los Alamos NM 87545, USA\n[1] E. H. Hall, Proc. Phys. Soc. Lond. 4, 325 (1880).\n[2] Y. Onose, N. Takeshita, C. Terakura, H. Takagi and Y.\nTokura, Phys. Rev. B 72, 224431 (2005).\n[3] R. Karplus and J. M. Luttinger, Phys. Rev. 95, 1154\n(1954).\n[4] T. Jungwirth, Qian Niu and A. H. MacDonald, Phys.\nRev. Lett. 88, 207208 (2002).\n[5] M. Onoda and N. Nagaosa, Phys. Rev. Lett. 90, 206601\n(2003).\n[6] J. Smit, Physica 21, 877 (1955); 24, 39 (1958).\n[7] L. Berger, Phys. Rev. 2, 4559 (1970).\n[8] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald and\nN. P. Ong, Rev. Mod. Phys. 82, 1539 (2010).\n[9] D. Xiao, M.-C. Chang and Q. Niu, Rev. Mod. Phys. 82,\n1959 (2010).\n[10] Hua Chen, Qian Niu and A. H. MacDonald, Phys. Rev.\nLett.112, 017205 (2014).\n[11] T. Asaba, S. M. Thomas, M. Curtis, J. D. Thompson,\nE. D. Bauer, and F. Ronning, Phys. Rev. B 101, 174415\n(2020).\n[12] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back\nand T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015).\n[13] G. Shirane, D. E. Cox, and S. J. Pickart, J. Appl. Phys.\n35, 954 (1964).\n[14] K. Tomiyasu, H. Hiraka, K. Ohoyama, and K. Yamada,\nJ. Phys. Soc. Jpn. 77, 124703 (2008).\n[15] K. Singh, A. Maignan, C. Simon, and C. Martin, App.\nPhys. Lett. 99, 172903 (2011).\n[16] A. Maignan, C. Martin, K. Singh, Ch. Simon, O. I. Lebe-\ndev, and S. Turner, J. Solid State Chem. 195, 41 (2012).\n[17] K. Tsuda, D. Morikawa, Y. Watanabe, S. Ohtani, and T.\nArima, Phys. Rev. B 81, 180102(R) (2010).\n[18] J. Bertinshaw, C. Ulrich, A. G¨ unther, F. Schrettle, M.\nWohlauer, S. Krohns, M. Reehuis, A. J. Studer, M.\nAvdeev, D. V. Quach, J. R. Groza, V. Tsurkan, A. Loidl\nand J. Deisenhofer, Sci. Rep. 4, 6079 (2014).\n[19] L. Lin, H. X. Zhu, X. M. Jiang, K. F. Wang, S. Dong, Z.\nB. Yan, Z. R. Yang, J. G. Wan, and J. M. Liu, Sci. Rep.\n4, 6530 (2014).\n[20] V. Tsurkan, O. Zaharko, F. Schrettle, C. Kant, J. Deisen -\nhofer, H. A. Krug von Nidda, V. Felea, P. Lemmens, J.\nR. Groza, D. V. Quach, F. Gozzo, and A. Loidl, Phys.\nRev B81, 184426 (2010).\n[21] V. Tsurkan, I. Fita, M. Baran, R. Puzniak, D.Samusi,\nR.Szymczak, H. Szymczak, S. Lkimm, M.Kliemm, S.\nHonn, and R. Tidecks, J. Appl. Phys. 90, 875 (2001).\n[22] A. P.Ramirez, R.J. Cava, andJ. Krajewski, Nature 386,\n156 (1997).\n[23] B. I. Min, S. S. Abik, H. C. Choi, S. K. Kwon, and J. S.\nKang, New Jour. Phys. 10, 055014 (2008).\n[24] G. J. Snyder, T. Caillat, and J. P. Fleurial, Phys. Rev.\nB62, 10185 (2000).\n[25] H. N. Ok, and C. S. Lee, Phys. Rev. B 33, 581 (1986).5\n[26] J. S. Kang, G. Kim, H. J. Lee, H. S. Kim, D. H. Kim, S.\nW. Han, S. J. Kim, C. S. Kim, H. Lee, J. Y. Kim, and\nB. I. Min, J. Appl. Phys. 103, 07D717 (2008).\n[27] C. S. Yadav, S. K. Pandey, and P. L. Paulose, arXiv:\n1904.06661.\n[28] Yu Liu, R. J. Koch, Zhixiang Hu, Niraj Aryal, Eli Stavit-\nski, Xiao Tong, Klaus Attenkofer, E. S. Bozin, Weiguo\nYin, and C. Petrovic, Phys. Rev. B 102, 085158 (2020).\n[29] J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev.\nLett.77, 3865 (1996).\n[30] G. Kresse and J. Furthm¨ uller, Phys. Rev. B 54, 11169\n(1996).\n[31] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza and D.\nVanderbilt, Rev. Mod. Phys. 84, 1419 (2012).\n[32] R. D. Barnard, Thermoelectricity in Metals and Alloys\n(Taylor & Francis, London, 1972).\n[33] K. Behnia, D. Jaccard and J. Flouquet, J. Phys.: Con-\ndens. Matter. 16, 5187 (2004).\n[34] K. Ueda, and T. Moriya, J. Phys. Soc. Jpn., 39, 605\n(1975).\n[35] A. Rosch, Phys. Rev. Lett. 82, 4280 (1999).\n[36] Q. Wang, S. S. Sun, X. Zhang, F. Pang, and H. C. Lei,\nPhys. Rev. B 94, 075135 (2016).\n[37] J. Yan, X. Luo, G. T. Lin, F. C. Chen, J. J. Gao, Y.\nSun, L. Hu, P. Tong, W. H. Song, Z. G. Sheng, W. J. Lu,X. B. Zhu, and Y. P. Sun, Europhys. Lett. 124, 67005\n(2018).\n[38] Y. H. Wang, C. Xian, J. Wang, B. J. Liu, L. S. Ling, L.\nZhang, L. Cao, Z. Qu, and Y. M. Xiong, Phys. Rev. B\n96, 134428 (2017).\n[39] S. Onoda, N. Sugimoto, and N. Nagaosa, Phys. Rev. B\n77, 165103 (2008).\n[40] O. Gunnarson, M. Calandra and J. E. Han, Rev. Mod.\nPhys.75, 1085 (2003).\n[41] S.Onoda, N.Sugimoto, andN.Nagaosa, Phys.Rev.Lett.\n97, 126602 (2006).\n[42] P. Nozi` eres and C. Lewiner, J. Phys. (Paris) 34, 901\n(1973).\n[43] J. P. Jan, and H. M. Gijsman, Physica 18, 339 (1952).\n[44] P. N. Dheer, Phys. Rev. 156, 637 (1967).\n[45] K. Ohgushi, S. Murakami and N. Nagaosa, Phys. Rev. B\n62, R6065 (2000).\n[46] R. Shindou and N. Nagaosa, Phys. Rev. Lett. 87, 116801\n(2001).\n[47] G. Tatara and H. Kawakamura, J. Phys. Soc. Jpn. 71,\n2613 (2002).\n[48] S. Gao, M. Hirschberger, O. Zaharko, T. Nakajima, T.\nKurumaji, A. Kikkawa, J. Shiogai, A. Tsukazaki, S.\nKimura, S. Awaji, Y. Taguchi, T. H. Arima and Y.\nTokura, Phys. Rev. B 100, 241115(R) (2019)." }, { "title": "1712.09973v1.Charge_ordering_and_ferrimagnetism_in_the_strongly_correlated__β__V__2_PO__5__single_crystal.pdf", "content": "Charge ordering and ferrimagnetism in the strongly correlated \f-V2PO 5single crystal\nJie Xing,1Huibo Cao,2Arpita Paul,3Chaowei Hu,4Hsin-Hua Wang,4Yongkang Luo,4\nRaj Chaklashiya,4Jared M. Allred,5Stuart Brown,4Turan Birol,3and Ni Ni1,\u0003\n1Department of Physics and Astronomy and California NanoSystems Institute,\nUniversity of California, Los Angeles, CA 90095, USA\n2Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA\n3Department of chemical engineering and materials science,University of Minnesota, MN 55455, USA\n4Department of Physics and Astronomy, University of California, Los Angeles, CA 90095, USA\n5Department of Chemistry and Biochemistry, University of Alabama, Tuscaloosa, AL 35487, USA\nA combined study of transport, thermodynamic, neutron di\u000braction, nuclear magnetic resonance\nmeasurements and \frst principles calculation were performed for \f-V2PO 5single crystal. It was\nshown to be a semiconductor with a band gap of 0.48 eV, undergoing a charge ordering (unusual\nV2+and V3+) phase transition accompanied by a tetragonal to monoclinic structural distortion at\n610 K and a paramagnetic to ferrimagnetic phase transition at 128 K with a propagation vector\nofk= 0. The easy axis is in the monoclinic acplane pointing 47(9)\u000eaway from the monoclinic a\naxis. This collinear ferrimagnetic structure and anisotropic isothermal magnetization measurements\nsuggest weak magnetic anisotropy in this compound. The \frst principles calculations indicate that\nthe intra-chain interactions in the face-sharing VO 6chains dominate the magnetic hamiltonian and\nidentify the \u0000+\n5normal mode of the lattice vibration to be responsible for the charge ordering and\nthus the structural phase transition.\nI. INTRODUCTION\nCharge ordering(CO), the long-range ordering of tran-\nsition metal ions with di\u000berent oxidization states, is a\nprominent feature in mixed valent 3 dtransition metal\noxides [1]. Due to the strong Coulomb interaction in the\ncharge ordering state, a high symmetry to low symmetry\nstructural distortion can occur, accompanied with the\nsudden enhancement in the electrical resistivity arising\nfrom the charge localization. The competition between\nthis charge disproportionation and the exchange inter-\nactions among magnetic transition metal ions has led to\nemergent phenomena, such as colossal magnetoresistance\nin RE 1\u0000xAxMnO 3(RE = rare earth, A = alkaline earth)\n[2, 3], superconductivity in \f-Ag 0:33V2O5[4], etc.\nThe vanadium phosphorus oxide system (V-P-O) has\ndistinct structural stacking and variable valences of\nvanadium, providing a great avenue to investigate the\nstructure-property relationship, enriching our under-\nstanding on the competition of CO and various exchange\ninteraction. The fundamental building blocks of the V-\nP-O system consist of VO 6octahedra or VO 4tetrahedra\nlinked by PO 4tetrahedra with valence P5+. Rich 3d\nvanadium magnetism and valences have been observed.\nFor example, VPO 4with V3+ions, containing one di-\nmensional chains of edge-sharing VO 6octahedra, under-\ngoes an incommensurate antiferromagnetic (AFM) phase\ntransition at 26 K and then a commensurate AFM phase\ntransition at 10.3 K [5]. \u000b-VO(PO 3)2with one dimen-\nsional chains of corner-sharing VO 6octahedra, is AFM\nat 1.9 K with valence V4+[6]. Mixed valence V3+and\nV4+antiferromagnetically couple together below 5 K in\n\u0003Corresponding author: nini@physics.ucla.eduV2(VO)(P 2O7)2, where segments of edge-sharing VO 4\ntetrahedra and VO 6octahedra exist [7]. Alternating V4+\nspin-chain model can be used to describe the magnetism\nin (VO) 2P2O7with corner and edge-sharing VO 6octa-\nhedra ladders [8{12].\nIn this article, we investigated \f-V2PO 5. In the tetrag-\nonal phase in Ref. [13] (Fig. 3(b)), it contains chains of\nface-sharing VO 6octahedra linked by PO 4tetrahedra.\nThese chains are stacked in layers along the caxis, run-\nning alternately along the aorbaxis in the adjacent\nlayers (Fig. 3b). This material is intriguing in three as-\npects. Firstly, the valence analysis with P5+and O2\u0000in-\ndicates remarkably low valence V2:5+in this compound,\nwhich may suggest possible CO of V3+and very uncom-\nmon valence V2+[14]. Secondly, the face-sharing VO 6\noctahedra in the building block is very rare for vana-\ndium oxides, implying unusually strong intra-chain in-\nteraction between V ions. What's more, although the\nparallel chains in each layer do not share any oxygen,\nthe perpendicularly-running chains in the neighboring\nlayers are corner-sharing. As a result, the other im-\nportant magnetic interaction is the inter-chain interac-\ntion between the corner-sharing V ions in neighboring\nlayers. Thirdly, a recent \frst-principles calculation sug-\ngested that the \f-V2PO 5is a ferromagnetic (FM) topo-\nlogical Weyl and node-line semimetal without any trivial\nband at the Fermi level [15], being a great material plat-\nform to study the emergent phenomena in FM topological\nsemimetals.\nDespite of these remarkable aspects we discussed\nabove, neither physical properties nor possible structural\ndistortion has been investigated for \f-V2PO 5, therefore,\nwe performed a combined study of the single crystalline\n\f-V2PO 5by x-ray and neutron di\u000braction as well as\nNMR, transport and thermodynamic measurements. We\ndiscovered that \f-V2PO 5is a semiconductor with a bandarXiv:1712.09973v1 [cond-mat.str-el] 28 Dec 20172\ngap of 0.48 eV. Upon cooling, a charge ordering phase\ntransition accompanied with a tetragonal to monoclinic\nstructural phase transition occurs at 610 K, followed by\na long range ferrimagnetic phase transition below 128 K.\nII. EXPERIMENTAL METHODS\nPrecursor\f-V2PO 5powder was made by solid state\nreaction. V 2O5powder and phosphorus chunks were\nweighed according to the stoichiometric ratio 1 : 1 and\nsealed in a quartz tube under vacuum. The ampule was\nslowly heated up to 600\u000eC and dwelled for 2 hours, and\nthen was increased to 1000\u000eC and stayed for 2 days be-\nfore it was quenched in water. The resultant \f-V2PO 5\n(\u00182 g) powder and iodine \rakes (10 mg / cm3) were\nloaded into a 15-cm long quartz tube and sealed under\nvacuum. Single crystals of \f-V2PO 5were then grown\nby chemical vapor transport method [13]. The hot end\nwas set at 1000\u000eC and the cold end was set at 900\u000eC.\nAfter two weeks, quite a few sizable three dimensional\nsingle crystals (\u00184 mm\u00024mm\u00022mm) were found at the\ncold end. The inset of Fig. 1(a) shows a \f-V2PO 5single\ncrystal against 1 mm scale.\nThroughout the paper, abplane is the plane where\nchains locate in. The (hkl) Tmeans the peak indexed\nin the tetragonal structure while (hkl) Mmeans the peak\nindexed in the monoclinic structure.\nMagnetic properties were measured in a Quantum\nDesign (QD) Magnetic Properties Measurement System\n(MPMS3). A single crystal around 20 mg with a pol-\nishedabsurface and a single crystal with as-grown (011)\nsurface were used. Temperature dependent heat capac-\nity was measured in a QD Dynocool Physical Properties\nMeasurement System (Dynoccol PPMS) using the relax-\nation technique at zero \feld. To enhance the thermal\ncontact and lower the measurement time, the \f-V2PO 5\nsingle crystal was ground into powder and then mixed\nwith silver powder according to the mass ratio of 1 : 1\n. The heat capacity of \f-V2PO 5was then obtained by\nsubtracting the heat capacity of silver [16]. Below 200\nK, the two wire ETO method was used for the electric\nresistivity measurement in PPMS. From 200 K to 400 K,\nthe electric resistivity was measured with standard four-\npoint method while above 400 K, it was measured in a\nhomemade high temperature resistivity probe.\nSingle crystal neutron di\u000braction was performed at\nthe HB-3A four-circle di\u000bractometer equipped with a\n2D detector at the High Flux Isotope Reactor(HFIR) at\nORNL. Neutron wavelength of 1.546 \u0017A was used from a\nbent perfect Si-220 monochromator [17]. The pyrolytic\ngraphite (PG) \flter was used before the sample to re-\nduce the half- \u0015neutrons. Representational analysis with\nSARAh [18] was run to search for the possible magnetic\nsymmetries. The nuclear and magnetic structure re\fne-\nments were carried out with the FullProf Suite[19]. Pow-\nder X-ray di\u000braction measurements were performed using\na PANalytical Empyrean di\u000bractometer (Cu K \u000bradia-tion). Using the Fullprof suit[19], Rietveld re\fnement\nwas carried out to re\fne the powder X-ray di\u000braction\ndata with the crystal structure determined by single crys-\ntal neutron di\u000braction.\nNuclear magnetic resonance (NMR) measurement was\ndone under a \fxed magnetic \feld of approximately 8.5 T,\napplied along the direction perpendicular to the abplane,\nwhere the chains locate in. The spectra were collected\nby performing an optimized \u0019=2-\u001c-\u0019spin-echo pulse se-\nquence. The spin-lattice relaxation time T1was mea-\nsured by integration of the phase corrected real part of\nthe spin echo using the saturation-recovery technique cite\nand spin echo decay time T2was measured by altering \u001c\nin the sequence. Spin-lattice relaxation time T1is ob-\ntained by the magnetization recovery \ftting to a single\nexponential form.\nFirst principles Density Functional Theory calcula-\ntions were performed to compare the energies of di\u000berent\nmagnetic con\fgurations. We used PAW as implemented\nin VASP with PBEsol exchange correlation functional\n[20{22]. A 8\u00028\u00028 k-point grid and energy cut o\u000b of\n500 eV ensures convergence in the primitive cell with 4\nformula units.\nIII. EXPERIMENTAL RESULTS\nA. Magnetic, transport and thermodynamic\nproperties\nFigure 1 (a)-(c) show the anisotropic magnetic prop-\nerties of\f-V2PO 5. Figure 1(a) presents the temperature\ndependent M=H taken atH= 1 kOe from 2 K to 250\nK in zero-\feld-cooled (ZFC) warming and \feld-cooled\nmode with Hparallel and perpendicular to the abplane.\nThe sharp upturns of the curves and the bifurcation in\nZFC and FC data for both directions below TIindicate\nthe existence of ferromagnetic component. The smooth\nZFC and FC curves suggest no other magnetic transi-\ntion below TI. Comparing with the other V-O-P mate-\nrials, the magnetic transition temperature is quite high\n[5{7], suggesting strong exchange interactions. Figure 1.\n(b) shows the temperature dependent M=H (blue) and\nH=M (black) measured from 300 K to 1000 K with H\n// (0 1 1) TatH= 10 kOe. Firstly, we see a subtle but\ndiscernable enhancement in M=H at the characteristic\ntemperature TII= 610 K, suggesting a possible phase\ntransition here. Secondly, upon cooling, linear Curie-\nWeiss behavior can be clearly seen from 1000 K to 500 K\ninH=M . By \ftting H=M from 1000 K to 500 K using the\nCurie-Weiss formula H=M =C=(T\u0000\u0012cw), whereCis the\nCurie constant and \u0012cwis the Weiss temperature, we ob-\ntained\u0016eff= 3.7(2)\u0016B/V and\u0012cw=\u0000900 K. The \u0016eff\nis larger than the one of V2:5+but comparable to the\none ofV2+. The large negative \u0012cwwithj\u0012cw=TIj\u00187.2,\nsuggests strong antiferromagnetic interaction. Thirdly,\ntheH=M fromTIto 500 K shows a crossover concave\nbehavior with temperature.3\nFIG. 1. (a) ZFC and FC M=H vs.TunderH= 1 kOe with H==ab andH?abfrom 2 K to 250 K. Inset: picture of \f-V2O5P\nsingle crystal against 1-mm scale. (b) H=M andM=H vs.TunderH=10 kOe with H==(011) T2 K to 1000 K. The red line\nis the Curie-Weiss \ft. (c) Isothermal M(H) curves at 2 K with H==ab andH?ab. Inset:M(H) curves at 300 K and 800\nK along with H==(011) T. (d) Speci\fc heat Cpvs.Tfrom 2 K to 200 K. The red line is the \ftting curve by Debye model.\nInset: Temperature dependence of magnetic entropy. (e) Resistivity \u001avs.Tfrom 130 K to 760 K. The red line emphasizes\nthe transition at TII. Inset:\u001avs. 1=T. Red line: the \ftting curve using the thermal excitation model. (f) The Arrott plot at\nvarious temperatures from 122 K to 137 K.\nFigure 1 (c) shows the anisotropic \feld dependent mag-\nnetizationM(H) taken at 2 K with H==ab andH?ab.\nThe crystal orientation was determined by x-ray di\u000brac-\ntion (Fig. S1) [23]. Clear hysteresis can be observed\nin both directions, con\frming the existence of ferromag-\nnetic component. Both curves show very similar shape\nand magnitude, suggesting weak magnetic anisotropy in\nthis system. The coercive \felds are around 1.3 kOe for\nboth. The remanent moments are 0.22 \u0016B/V forH?ab\nand 0.25\u0016B/V forH==ab and the saturation moments\nare 0.27\u0016B/V forH?aband 0.31\u0016B/V forH==ab ,\nwhich are so much smaller than the saturation moment\nof V2+(d3) and V3+(d2) ions, suggesting that instead\nof ferromagnetism, this material is likely ferrimagnetic or\ncanted antiferromagnetic below TI. The inset of Fig. 1(c)\nshows theM(H) curves taken at 300 K and 800 K with\nH// (0 1 1) T. Both curves are linear with the applied\nmagnetic \feld without hysteresis. The Arrott plot ( M2\nvs.H=M ) has been widely used to determine the fer-\nromagnetic phase transition temperature [24, 25], where\nthe curve of M2vs.H=M passes through the origin of\nthe plot at the transition temperature. To determine the\nvalue ofTI, isothermal M(H) curves are measured from\n122 K to 137 K. The Arrott plot are calculated and shown\nin Fig. 1(f), which suggests that TI\u0018128 K.\nFigure 1 (d) shows the temperature dependent Cp=T\ndata (blue) taken from 2 K to 200 K. A heat capacityanomaly featuring a second order phase transition ap-\npears around 128 K, accompanying with the magnetic\nphase transition observed in Fig. 1 (a). To estimate the\nmagnetic entropy, Debye model is used to \ft the heat\ncapacity to provide the non-magnetic background. The\n\ftted curve is shown in red in Fig. 1(d) and the \ftted\nDebye temperature is \u0002 D= 620 K. By subtracting the\nnon-magnetic background from \ftting, we obtained the\nmagnetic entropy as SM= 5.6 J/mol V-K2. This value is\nsigni\fcantly smaller than RLn4 of V2+andRLn3 of V3+,\nbut rather approximate RLn2. This may be caused by\nthe strong V-O covalency which lowers the moment size\nof V or a result of the entropy release above 128 K due\nto the chain structure and strong intrachain interaction\n[26, 27].\nFigure 1(e) shows the resistivity ( \u001a) of the\f-V2PO 5\nsingle crystals vs. temperature from 130 K up to 760 K.\nInstead of the semi-metal suggested by the theoretical\nprediction [15], it is a semiconductor. A semiconductor\nto semiconductor phase transition is discernable at 610\nK, which con\frms the possible phase transition at TII\nsuggested by the subtle susceptibility increase shown in\nFig. 1 (b). \u001avs. reciprocal temperature is plotted in\nthe inset of Fig. 1(d). By \ftting the data between 280\nK to 130 K with the thermal excitation model \u001a(T) =\n\u001a(0)exp(Eg=2kBT), the estimated gap size of \f-V2PO 5\nis 0.48 eV. The gap value is similar to 0.45-0.57 eV of the4\nVanadium phosphate glass [28].\nB. Structural phase transition and charge ordering\nBased on the transport, magnetic and heat capacity\nmeasurements, we have shown that \f-V2PO 5has one\nphase transition at 610 K and the other magnetic phase\ntransition at 128 K. To investigate the nature of these\ntwo phase transitions, single crystal neutron di\u000braction\nand NMR measurements are performed, which are sum-\nmarized in Fig. 2.\nFigure 2 (a) and (c) show the order parameter plot\nof the (1 1 4) Tneutron peak up to 650 K and (1 1 0) T\nneutron peak up to 450 K, respectively and Fig. 2 (b)\npresents the rocking curve scan for the (1 1 4) Tpeak. It is\ntwinned structure below 610 K, so we keep the tetragonal\nindex for convenience. The order parameter plot of the\nrelative stronger peak (1 1 4) T(Fig. 2(a)) indicates two\nphase transitions occurring at 610 K and 128 K, respec-\ntively, which is consistent with the order parameter plot\nof the (1 1 0) Tpeak (Fig. 2(c)) and (1 0 1) Tpeak (Fig.\nS2) [23]. The full data were collected at 4.5 K, 300 K,\nand 650 K to cover all three phase regions. At 650 K, the\ndata can be well \ftted in I41/amd symmetry (Table I).\nSince both (1 1 0) Tand (1 1 4) Tpeaks are symmetry dis-\nallowed re\rections in the tetragonal I41/amd structure,\nthe fact that we observed these peaks at room temper-\nature (Fig. 2(b)) suggests possible structural/magnetic\nphase transitions at 610 K.\nTo identify if the phase between 128 K and 610 K has\na magnetic component, phosphorus-31 NMR (31P-NMR)\nmeasurements were carried out. These are summarized\nin Fig. 2(e)-(f). In Fig. 2(d), we report the31P-NMR\nspectra at various temperatures from 170 K to 300 K.\n31P nuclear spin I= 1=2, and all sites are equivalent\nforB?ab. At 300 K, the spectra shift from the Lar-\nmor frequency by Ks= 0:493\u00060:002%, where the mean\nand uncertainty were calculated using the gaussian \ft-\nting. This value is on the same order of another VPO\nsample with V3+and about twice as much as that with\nV4+[29, 30], where a similar shift and broadening were\nobserved. The observations are interpreted as evidence\nfor no long range magnetic ordering in the intermediate,\ncharge-ordered phase. Below 190 K, a minor absorption\npeak at 146.4 MHz is resolved. Since it accounts for\nonly 3% of the total spin intensities, we expect the signal\nto be extrinsic due to the sites at twin boundaries and\nexclude it in our analysis. From the slope of K- \u001fplot\nshown in Fig.2 (e), we estimate the hyper\fne coupling\nconstant to be Acc= (10:23\u00060:87) kOe/\u0016B, which is\ntransferred from the unpaired electrons from the second\nnearest neighbors of P [31]. Figure 2(f) shows the tem-\nperature dependence of the spin-lattice relaxation time\nT1and spin-spin relaxation time T2. The relatively short\nand constant T1is consistent with a paramagnetic phase\nfrom 150 k to 300 K [32]. Meanwhile, T2starts drop-\nping rapidly below 210 K as the system approaches the\nFIG. 2. (a) The (1 1 4) Tneutron peak intensity vs. T. (b)\nThe (1 1 4) Tneutron peak intensity vs. !. (c) The (1 1 0) T\nneutron peak intensity vs. T. (d) P NMR frequency spectra.\nThe spectra are conserved after corrections for T 2. (e) Knight\nshift K and the peak width \u0001F obtained from (a) vs. mag-\nnetic susceptibility \u001f. Dashed line shows a linear \fttings of\nK. (f) Spin-lattice relaxation time T1and spin-spin relaxation\ntimeT2vs. T.\n128 K transition. This behavior is associated with slow\nlongitudinal \ructuations, which are likely related with\nthe onset of the 128 K magnetic phase transition. This\nis also consistent with the broadening observed in Fig.\n2(d) which appears as the magnetic correlation develops\nupon cooling. However, since 1/ T2is on the order of a\nfew to hundreds of KHz, we conclude that most of our\nbroadening, which is on the order of several MHz, is from\nthe inhomogeneous internal \feld.\nSince NMR shows that the phase transition at 610 K\nis not of magnetic origin, the phase transition at 610\nK should be a structural distortion. By including four\ntwinned structure domains, the room temperature neu-\ntron data can be well \ftted with the C2/cmonoclinic\nsymmetry, suggesting a tetragonal to monoclinic phase\ntransition at 610 K. Using this crystal structure, we re-\n\fned our powder X-ray di\u000braction taken at room tem-\nperature and obtained a very good \ft as shown in Fig.5\nFIG. 3. (a) The experimental and re\fned powder X-ray\ndi\u000braction patterns for \f-V2O5P at 300 K. Black: experi-\nmental pattern. Red: re\fned pattern. green: the di\u000berence\nbetween the experimental and re\fned patterns. Black ticks:\nthe Bragg peak positions in the monoclinic structure. Inset:\nenlarged view from 55.5\u000eto 58\u000e. (b)(c): The crystal structure\nof\f-V2O5P at 650 K (b) and 300 K (c).\n3(a). The detailed crystal structures at 4.5 K, 300 K\nand 650 K are summarized in Table I. The high temper-\nature tetragonal and low temperature monoclinic struc-\ntures are visualized in Fig. 3(b) and (c), respectively. In\nthe monoclinic phase, the monoclinic caxis is 121.45\u000e\nfrom theabplane which is the plane where the chains\nsit in. The unique V site in the tetragonal structure sep-\narates into V1 and V2 sites with the V1 atoms and V2\natoms alternately locating along each chain direction (V1\nand V2 sites are labeled in Fig. 3(c)), as a result, the av-\nerage bond length of VO 6octahedra on V1 site increases\nwhile that on V2 site decreases. Bond-valence analysis\nof the monoclinic crystal structure assigns charges of 2.0\nand 2.9 toV1andV2respectively, which is a smoking gun\nproof of the charge order [33, 34].\nC. Ferrimagnetic structure below 128 K\nThe magnetic order onsets at 128 K while the charge\norder continues to develop below the magnetic transi-TABLE I. The crystal structure of the \f-V2OPO 4phase at\n4.5 K, 300 K and 650 K, respectively.\n\f-V2PO 5at 4.5 K monoclinic C2/c\na= 7.563 \u0017A b=7.563 \u0017A c=7.235 \u0017A\f= 121.51\u000e\nRF2=0.0691 wRF2=0.0841 R F=0.044 \u001f2=19.2\nsite x/a y/b z/c\nV1 0 1/2 0\nV2 1/4 1/4 0\nO1 0.066(3) 0.749(2) 0.632(2)\nO2 0.317(2) 0.496(2) 0.598(1)\nO3 0 0.652(2) 1/4\nP 0 0.121(2) 1/4\n\f-V2PO 5at 300 K monoclinic C2/c\na= 7.570 \u0017A b=7.570 \u0017A c=7.232 \u0017A\f= 121.56\u000e\nRF2=0.0734 wRF2=0.0956 R F=0.0473 \u001f2=25.1\nsite x/a y/b z/c\nV1 0 1/2 0\nV2 1/4 1/4 0\nO1 0.065(3) 0.749(2) 0.630(2)\nO2 0.318(2) 0.496(2) 0.599(1)\nO3 0 0.652(2) 1/4\nP 0 0.119(3) 1/4\n\f-V2PO 5at 650K Tetragonal I41/amd\na= 5.357(2) \u0017A b= 5.357(2) \u0017A c=12.373(4) \u0017A\nRF2=0.0707 wRF2=0.0888 R F=0.0421 \u001f2=5.12\nsite x/a y/b z/c\nV1 1/4 1/4 1/4\nP1 0 3/4 1/8\nO1 0 -0.012(1) 0.193(4)\nO2 0 3/4 5/8\nFIG. 4. (a)-(c): Three magnetic structure models showing\nferrimagnetism with di\u000berent easy axis. Model (a) is collinear\nsuggesting weak magnetic anisotropy while Models (b) and\n(c) are noncollinear indicating strong magnetic anisotropy.\nModel (a) is the magnetic structure of \f-V2PO 5.6\ntion. Since no observed sharp change can be determined\nby the structure re\fnement at 4 K (see Table I), no fur-\nther structural phase transition below 128 K is discern-\nable. The magnetic propagation vector is k=0, which\nmeans that the magnetic scattering signal appears on\ntop of the nuclear Bragg peaks. To determine the mag-\nnetic structure more precisely, the magnetic signals were\nextracted by subtracting the data measured just above\n128 K from that at 4 K. During the procedure, to se-\nlect peaks which are insensitive to the thermal displace-\nments and charge ordering, we compared the data mea-\nsured at 300 K and 450 K and only selected a peak if\nthe change of its intensity is much smaller than the ex-\ntracted magnetic intensity. The selected re\rections are\nlisted in Table SI [23]. Since (1 1 0) T, (1 1 4) T, (1 0\n1)T, (0 0 2) Tand (0 0 4) Tpeaks were measured with a\nlong counting time and also tracked upon warming, they\nare highly reliable as indicated by their small error bars\nshown in the Table SI. We then performed the represen-\ntational analysis that determines the symmetry-allowed\nmagnetic structures for a second-order magnetic tran-\nsition. It yielded two magnetic symmetries, C2=cand\nC20=c0. Only the C20=c0can \ft our data (see Table SI).\nThe obtained magnetic structure is ferrimagnetic. Spins\non all V2+(V1 sites) atoms are parallel and so do the\nspins on all V3+(V2 sites) atoms while these two spin\nsublattices are antiparallel to each other. Since it is un-\nlikely for the V2+(d3) to be in a low spin state due to\nthe longer V-O bond length on V1 site and thus weaker\ncrystal electric \feld, the moment MV1> M V2. Figure\n4 shows the ferrimagnetic structure with three possible\neasy axis assignments where MV1> M V2. The calcu-\nlated peak intensity and goodness of \ft are summarized\nin Table SI. The model in the left panel of Fig. 4(a)\nis our pick for the \f-V2PO 5which gives the best \ft of\nthe data as shown in the right panel of Fig. 4(a). The\nMV1= 1:4(1)\u0016BandMV2= 1:2(1)\u0016B. The easy axis is\ninacMplane and 47(9) degrees away from aMtowards\ncM. The magnetic structure is collinear, suggesting weak\nmagnetic anisotropy. This is indeed consistent with the\nanisotropic M(H) measurements shown in Fig. 1(c). The\nother two models shown in Fig. 4(b) and (c) are with the\neasy axis along the chain direction (Fig. 4(b)) or perpen-\ndicular to the chain direction on the abplane (Fig. 4(c)).\nThese two magnetic models are non-collinear with strong\nmagnetic anisotropy. Since the goodness of \ft for these\ntwo latter models are poor, they are not the right mag-\nnetic structure for \f-V2PO 5.\nIV. DISCUSSION\nIt is of particular interest to ask if the charge order is\nsolely responsible for the reduction in the crystal symme-\ntry, or rather if it is a secondary order parameter to some\nother electronic phase transition. In order to preclude\nthis possibility and elucidate the nature of the charge or-\ndering transition at 610 K, we performed a group theo-TABLE II. Energies of di\u000berent magnetic con\fgurations from\n\frst principle.\nIntra-chain Inter-chain Energy (meV/f.u.)\nFerrimagnetic Ferromagnetic 0\nFerrimagnetic Antiferromagnetic 9\nFerromagnetic Ferromagnetic 60\nFerromagnetic Antiferromagnetic 71\nretical analysis of the lattice distortion using the Isotropy\nSoftware Suite.[35] The distortion from the high temper-\nature tetragonal structure ( I41=amd ) to the low temper-\nature monoclinic structure ( C2=c) can be caused by two\nnormal modes of lattice vibration, \u0000+\n5and \u0000+\n4. \u0000+\n5re-\nduces the symmetry from I41=amd toC2=c, and is the\nonly candidate for the primary structural order parame-\nter, whereas \u0000+\n4by itself reduces the symmetry to Fddd .\nC2cis a subgroup of Fddd , and as a result, \u0000+\n4is most\nlikely a secondary order parameter that is not important\nin the energetics of the phase transition. At the same\ntime, the charge order itself, which is a di\u000berentiation\nof the neighboring V ions in the same chain, transforms\nas the \u0000+\n5irreducible representation does for the high\nsymmetry structure. Figure 5(a) shows the displacement\nof the oxygen atoms in the VO 6face-sharing chain due\nto the \u0000+\n5normal mode. This mode breaks the symme-\ntry between the V ions that are symmetry equivalent at\nI41=amd , and decreases the V-O bond length for V2while\nincreasing it for V1(Fig. 5(b)) as expected in a charge\nordering transition. We therefore conclude that to re-\nduce the symmetry to the monoclinic phase, the charge\norder, by itself, is su\u000ecient and no other magnetic or\nelectronic mechanisms are necessary. This is consistent\nwith the fact that the TCOis almost 5 times of the Tmag.\nWe also note that \u0000+\n5is a Raman active mode, and as a\nresult, signature of the charge ordering transition should\nbe visible in the Raman spectrum of \f-V2PO 5.\nWe performed the \frst principles calculation using\nDFT+U with U = 4 eV to correct for the underestima-\ntion of the on-site coulomb interaction on the V ion [36].\nOur DFT calculations predict magnetic moments of 2.6\n\u0016Band 1.8\u0016Brespectively inside the V1andV2spheres,\nbut the band structure (Fig. 5(c)) shows no partially\n\flled bands. This signals strong hybridization between\ntheVand theOions. DFT gives a magnetic moment\n0.5\u0016BperVincluding the interstitials. This is a strong\noverestimation compared to the experimental value. The\nreason of this is likely the DFT+U's tendency to overes-\ntimate the ordered moments when there are dynamical\n\ructuations present, and it is possible that a more ad-\nvanced \frst principles method (such as the Dynamical\nMean Field Theory) can reproduce the experimentally\nobserved value of the local moments.\nTo understand the magnetic order, we calculated the\nenergies of phases with di\u000berent magnetic orders from\nDFT, as listed in Table II. The lowest energy phase is\npredicted to have ferrimagnetic intra-chain order, where7\nFIG. 5. (a) A sketch of the Oxygen anion displacements due\nto the \u0000+\n5mode which is responsible for charge ordering. (b)\nThe average V\u0000Obond length as a function of the \u0000+\n5normal\nmode amplitude. (c) DFT band structure in the ferrimagnetic\nstate. Majority and minority spin bands are shown in red and\nblue respectively.\nthe moments of the neighboring V1andV2ions on the\nsame chain are aligned anti-parallel, and ferromagnetic\ninter-chain order, so that, for example, the magnetic mo-\nments of all V1ions are parallel. This observation is in\nline with the experimental observation. The energy cost\nof having a magnetic phase where the di\u000berent chains\nhave antiparallel moments is bout 9-10 meV per formula\nunit, whereas the energy cost of having spins on the same\nchain parallel is 60 meV per formula unit. This suggests\nthat the intra-chain interactions between the V1andV2\nions in the face-sharing VO 6chains are the dominant\nterm in the magnetic hamiltonian.\nTo understand the crossover behavior in H/M shown\nin in Fig. 1(b), we calculated the energetics of di\u000ber-\nent magnetic phases in the high temperature tetrago-\nnal structure ( I41=amd ) with the same parameters (not\nshown), and found that similar couplings apply to mag-netic moments in that structure too. This explains the\ncross-over: Above the charge ordering temperature, the\nmagnetic moments on all V ions are equal, and its Curie-\nWeiss behaviour is that of an antiferromagnet, with a\nnegative Curie temperature. However, charge ordering\nmakes the moments unequal, and as a result, below 610\nK the Curie-Weiss behaviour is that of an ferrimagnet,\nwhich has a positive Curie temperature like a ferromag-\nnet.\nTo address if there is non-trivial topology in this com-\npound, the DFT band structure in Fig. 5(c) is calcu-\nlated in the ferrimagnetic ground state. Unlike the DFT\nband structure calculated in the ferromagnetic phase and\nthe tetragonal structure without U [15], there are no\nband crossing at the Fermi level. And more importantly,\nthe top of the valence and the bottom of the conduc-\ntion bands have opposite spin directions. This obser-\nvation precludes any possibility of topological phases in\n\f-V2PO 5.\nV. CONCLUSION\nIn conclusion, we have grown and characterized \f-\nV2PO 5single crystals. A tetragonal to monoclinic struc-\ntural phase transition at 610 K and a paramagnetic to\nferrimagnetic phase transition are revealed by transport,\nmagnetic, speci\fc heat, single crystal neutron di\u000braction\nand NMR measurements. Below 610 K, the single V site\nat the high temperature tetragonal phase distorts into\ntwo alternating V sites, leading to the increase of V-O\nbond length of one V site but the decrease of the other\nV site. Our \frst principles calculation shows that this\ndistortion was caused by the \u0000+\n5normal mode of lat-\ntice vibration. Accompanied with the distortion, charge\nordering of V2+and V3+is undoubtedly suggested by\nthe Bond-valence analysis. Below 128 K, the spins order\nparallel on each sublattice of V sites while the spins on\none sublattice order antiparallel to the other. With the\neasy axis being in the monoclinic acplane and 47(9)\u000e\naway from the monoclinic atowardscaxis., this gives a\ncollinear ferrimagnetic structure with the moment to be\n1.4(1)\u0016Bon V2+site and 1.2(1) \u0016Bon V3+site, suggest-\ning weak magnetic anisotropy and dominant role of the\nintra-chain V-V interaction in magnetism. No non-trivial\ntopology is suggested by our \frst principles calculation.\nACKNOWLEDGMENTS\nWork at UCLA (JX, CWH, RC, NN) was supported\nby NSF DMREF program under the award NSF DM-\nREF project DMREF-1629457. Work at ORNL HFIR\nwas sponsored by the Scienti\fc User Facilities Division,\nO\u000ece of Science, Basic Energy Sciences, U.S. Depart-\nment of Energy. Work at UMN was supported by NSF\nDMREF program under the award NSF DMREF project\nDMREF-1629260. Work at UCLA (HHW, YKL, SB) was8\nsupported by NSF DMR-1410343 and DMR-1709304.\nYKL would also like to thank the support from LANL\nLDRD program.\nNote : During the preparation of this manuscript, we\nnoticed a work on pollycrystalline \f-V2OPO 4was justaccepted but Journal of the american chemistry society\n(http://pubs.acs.org/doi/abs/10.1021/jacs.7b09441),\nwhich also revealed the charge ordering and simi-\nlar ferrimagnetism with di\u000berent easy axis in this\ncompound.\n[1] J. Att\feld. Charge ordering in transition metal oxides.\nSolid state sciences 8, 861 (2006).\n[2] E. Wollan, W. Koehler. Neutron Di\u000braction Study of\nthe Magnetic Properties of the Series of Perovskite-Type\nCompounds [(1-x)La,xCa]MnO 3.Phys. Rev. 100, 545\n(1955).\n[3] J. Goodenough. Theory of the Role of Covalence in\nthe Perovskite-Type Manganites [La,M(II)]MnO 3.Phys.\nRev.100, 564 (1955).\n[4] T. Yamauchi, M. Isobe, and Y. Ueda. Charge order\nand superconductivity in vanadium oxides. Solid State\nSciences 7, 874 (2005).\n[5] R. Glaum, M. Reehuis, N. St u\fer, U. Kaiser, and\nF. Reinauer. Neutron Di\u000braction Study of the Nuclear\nand Magnetic Structure of the CrVO 4Type Phosphates\nTiPO 4and VPO 4.J.Solid State Chem. 126, 15 (1996).\n[6] J. Kikuchi, N. Kurata, K. Motoya, T. Yamauchi,\nand Y. Ueda. Spin Di\u000busion in the S=1/2 Quasi\nOne-Dimensional Antiferromagnet \u000b-VO(PO 3)2via 31P\nNMR. J.Phys. Soc. Jpn.70, 2765 (2001).\n[7] J. Johnson, D. Johnston, H. King Jr., T. Halbert, J.\nBrody, and D. Goshorn. Structure and magnetic proper-\nties of V 2(VO)(P 2O7)2. A mixed-valence vanadium (III,\nIII, IV) pyrophosphate. Inorg. Chem. 27, 1646 (1988).\n[8] D. Johnston, J. Johnson, D. Goshorn, and A. Jacob-\nson. Magnetic susceptibility of (VO) 2P2O7: A one-\ndimensional spin-1/2 Heisenberg antiferromagnet with\na ladder spin con\fguration and a singlet ground state.\nPhys. Rev. B35, 219 (1987).\n[9] E. Dagotto, J. Riera, and D. Scalapino. Superconductiv-\nity in ladders and coupled planes. Phys. Rev. B45,5744\n(1992).\n[10] A. W. Garrett, S. E. Nagler, D. A. Tennant, B. C. Sales,\nand T. Barnes. Magnetic excitations in the S=1/2 alter-\nnating chain compound (VO) 2P2O7.Phys. Rev. Lett.79,\n745 (1997).\n[11] J. Kikuchi, K. Motoya, T. Yamauchi, and Y. Ueda. Coex-\nistence of double alternating antiferromagnetic chains in\n(VO) 2P2O7: NMR study. Phys. Rev. B60, 6731 (1999).\n[12] T. Yamauchi, Y. Narumi, J. Kikuchi, Y. Ueda, K. Tatani,\nT. C. Kobayashi, K. Kindo, and K. Motoya, Phys. Rev.\nLett. 83, 3729 (1999).\n[13] R. Glaum, and R. Gruehn. Synthese, Kristallstruktur\nund magnetisches Verhalten von V2PO5. Z.Kristallogr\n186, 91 (1989).\n[14] D. C. Johnston, Magnetic Susceptibility of Collinear and\nNoncollinear Heisenberg Antiferromagnets, Phy. Rev.\nLett., 109, 077201 (2012)\n[15] Y. Jin, R. Wang, Z. Chen, J. Zhao, Y. Zhao, and H.\nXu. Ferromagnetic Weyl semimetal phase in a tetragonal\nstructure. Phys. Rev. B96, 201102 (2017).\n[16] D. Smith, and F. Fickett. Low-temperature properties of\nsilver. J.Res. Natl. Inst. Stand. Technol. 100, 119 (1995).\n[17] B. Chakoumakos, H. Cao, F. Ye, A. Stoica, M. Popovici,M. Sundaram, W. Zhou, J. Hicks, G. Lynn and R. Riedel.\nFour-circle single-crystal neutron di\u000bractometer at the\nHigh Flux Isotope Reactor. J.Appl. Crystallogr. 44, 655\n(2011).\n[18] A. Wills. A new protocol for the determination of\nmagnetic structures using simulated annealing and rep-\nresentational analysis (SARA h).Phys. B:Condens.\nMatt. 276, 680 (2000), program available from\nwww.ccp14.ac.uk.\n[19] J. Rodriguez-Carvajal. Recent advances in magnetic\nstructure determination by neutron powder di\u000braction.\nPhys. B:Condens. Matt. 192, 55 (1993).\n[20] G. Kresse and J. Furthm uller. E\u000ecient iterative schemes\nfor ab initio total-energy calculations using a plane-wave\nbasis set. Phys. Rev. B54, 11169 (1996).\n[21] G. Kresse, and J. Hafner. Ab initio molecular dynamics\nfor open-shell transition metals. Phys. Rev. B48, 13115\n(1993).\n[22] J. Perdew, A. Ruzsinszky, G. Csonka, O. Vydrov, G.\nScuseria, L. Constantin, X. Zhou, and K. Burke. Restor-\ning the Density-Gradient Expansion for Exchange in\nSolids and Surfaces. Phys. Rev. Lett. 100, 136406 (2008).\n[23] See supplimentary material.\n[24] A. Arrott, and J. Noakes. Approximate equation of state\nfor nickel near its critical temperature. Phys. Rev. Lett.\n191,786 (1967).\n[25] M. Halder, S. M. Yusuf, M. D. Mukadam, and K.\nShashikala, Phys. Rev. B 81, 174402 (2010)\n[26] L. Jongh, and A. Miedema. Experiments on simple mag-\nnetic model systems. Advances inPhysics 23,1 (1974).\n[27] S. Nagata, P. Keesom, and S. Faile. Susceptibilities of the\nvanadium Magn\u0013 eli phases V nO2n\u00001at low temperature.\nPhys. Rev. B20, 2886 (1979).\n[28] M. Khan, R. Harani, M. Ahmed, and C. Hogarth. A\ncomparative study of the e\u000bects of rare-earth oxides on\nthe physical, optical, electrical and structural properties\nof vanadium phosphate glasses. J.Materials Science 20,\n2207 (1985).\n[29] M. Sananes, and A. Tuel. Study by 31 P NMR spin echo\nmapping of vanadium phosphorus oxide catalysts. Solid\nstate nuclear magnetic resonance 6, 157 (1996).\n[30] M. Sananes, A. Tuel, and J. C. Volta. A study by 31P\nNMR spin-echo mapping of VPO catalysts: I. charac-\nterization of the reference phases. J.Catalys. 145, 251\n(1994).\n[31] J. Li, M. Lashier, G. Schrader, B. Gerstein. Oxida-\ntion states of vanadium in V-P-O oxidation catalysts\n31P NMR by spin-echo mapping. Appl. Catalys. 73, 83\n(1991).\n[32] T. Moriya. Nuclear magnetic relaxation in antiferromag-\nnetics. Prog. Theor. Phys. 16, 23 (1956).\n[33] N. Brese, and M. O'kee\u000be. Bond-valence parameters\nfor solids. Acta Crystallographica Section B:Structural\nScience 47, 192 (1991).9\n[34] I. David Brown, Bond valence parameters,\nhttps://www.iucr.org/resources/data/datasets/bond-\nvalence-parameters\n[35] H. Stokes, D. Hatch and B. Campbell. Isotropy. Retrievedfrom stokes. byu. edu/isotropy. html(2007).\n[36] A. Liechtenstein, V. Anisimov and J. Zaanen. Density-\nfunctional theory and strong interactions: Orbital order-\ning in Mott-Hubbard insulators. Phys. Rev. B,52, R5467\n(1995)." }, { "title": "2308.07236v3.Temperature_Evolution_of_Magnon_Propagation_Length_in_Tm__3_Fe__5_O___12___Thin_Films__Roles_of_Magnetic_Anisotropy_and_Gilbert_Damping.pdf", "content": " \n1 \n Temperature Evolution of Magnon Propagation Length in \nTm 3Fe5O12 Thin F ilms: Roles of Magnetic Anisotropy and Gilbert \nDamping \n \nAmit Chanda1, Christian Holzmann2, Noah S chulz1, Aladin Ullrich2, Derick DeTellem1, \nManfred Albrecht2*, Miela J. Gross3, Caroline A. Ross3*, Dario A. Arena1, Manh -Huong \nPhan1, and Hari haran Srikanth1* \n1Department of Physics, University of South Florida, Tampa, Florida 33620, USA \n2Institute of Physics, University of Augsburg, 86159 Augsburg, Germany \n3Department of Materials Science and Engineering, Massachusetts Institute of Technology, \nCambridge, Massachusetts 02139, USA \n \n*Corresponding authors: manfred.albrecht@physik.uni -augsburg.de ; caross@mit.edu ; \nsharihar@usf.edu \n \nKeywords: Longitudinal spin Seebeck effect, Inverse spin Hall effect, Magnon propagation length, \nGilbert damping, Magnetic anisotropy , Rare -earth iron garnet \n \nABSTRACT \nThe magnon propagation length , 〈𝜉〉 of a ferro -/ferrimagnet (FM) is one of the key factors that \ncontrols the generation and propagation of thermally -driven magnonic spin current in FM/heavy \nmetal ( HM) bilayer based spincaloritronic devices . For the development of a complete physical \npicture of thermally -driven magnon transport in FM/HM bilayers over a wide temperature range, \n2 \n it is of utmost importance to understand the respective roles of temperature -dependent Gilbert \ndamping (𝛼) and effective magnetic anisotropy (𝐾𝑒𝑓𝑓) in controlling the temperature evolution of \n〈𝜉〉. Here, we report a comprehensive investigation of the temperature -dependent longitudinal spin \nSeebeck effect (LSSE), radio frequency transverse susceptibility, and broadband ferromagnetic \nresonance measurements on Tm 3Fe5O12 (TmIG )/Pt bilayers grown on different substrates. We \nobserve a remarkable drop in the LSSE voltage below 200 K independent of TmIG film thickness \nand substrate choice . This is attribute d to the noticeable increases in effective magnetic anisotropy \nfield, 𝐻𝐾𝑒𝑓𝑓 (∝𝐾𝑒𝑓𝑓) and 𝛼 that occur within the same temperature range. From the TmIG \nthickness dependence of the LSSE voltage, we determined the temperature dependence of 〈𝜉〉 and \nhighlighted its correlation with the temperature -dependent 𝐻𝐾𝑒𝑓𝑓 and 𝛼 in TmIG/Pt bilayers , which \nwill be beneficial for the development of rare -earth iron garnet -based efficient spincaloritronic \nnanodevices . \n \n \n \n \n \n \n \n \n \n3 \n 1. INTRODUCTION \nIn recent years, interface -engineered bilayer thin films have gained intense attention of the \nmaterials science community because of their multifunctionality and emergent physical properties \nranging from ferroelectricity1 and magnetism2 to spin -electronics3. Bilayers comprised of \ninsulating rare -earth iron garnet (REIG ) and heavy metal (HM) form the most appealing platform \nto generate, transmit , and detect pure spin currents in the field of spin -based -electronics4–6. The \ninterplay of damping and magnon propagation length ( 〈𝜉〉) of the REIG layer and spin-orbit \ncoupling (SOC) of the HM layer leads to a wide range of emergent spintronic phenomena in t his \nfascinating class of heterostructures , including the spin Hall effect7, spin -orbit torque8,9, spin-\npumping effect (SPE)10, and the longitudinal spin Seebeck effect (LSSE)11–13. The discovery of \nthe SSE14 instigated a new generation of spintronic nano devices facilitating electrical energy \nharvesting from renewable thermal energy wherein a magnonic spin current is thermally generated \nand electrically detected by applying a temperature gradient across a magnetic insulator (MI) /HM \nbilayer15. Unlike magnetostatic spin waves with millimeter -range propagation lengths, 〈𝜉〉 for \nthermally generated magnons is significantly smaller, a few hundreds of nanometers16. In the \nframework of an atomistic spin model based on linear spin -wave theory, it was theoretically \nshown17,18 that thermally generated magnons have a broad frequency (𝑓) distribution with \n𝑓𝑚𝑖𝑛𝑖𝑚𝑢𝑚=2𝐾𝑒𝑓𝑓[ℎ(1+𝛼2)] ⁄ and 𝑓𝑚𝑎𝑥𝑖𝑚𝑢𝑚 =4𝐾𝑒𝑓𝑓[ℎ(1+𝛼2)] ⁄ , where ℎ is the Planck \nconstant, 𝐾𝑒𝑓𝑓 is the effective magnetic anisotropy constant and 𝛼 is the Gilbert damping \nparameter. While the high -f magnons experience stronger damping, low -f magnons possess a very \nlow group velocity, and hence, the majority of the thermally generated magnons become damped \non shorter length -scales17,18. Therefore, only the subthermal magnons, i.e., the low-f magnons \n4 \n dominate the long -range thermo -spin transport19–21. Within this hypothesis, it was predicted that \n〈𝜉〉 is inversely proportional to both 𝛼 and √𝐾𝑒𝑓𝑓.17,18 \n \nY3Fe5O12 (YIG) has been a widely explored MI for generating and transmitting pure spin \ncurrents due to its ultra -low damping (𝛼 ≈ 10-4-10-5) and large 〈𝜉〉 (~100-200 nm) 11,13,17. This has \nled to a drastic increase in research over the last few decade s, aimed at enhancing the spin current \ninjection efficiency across the MI/HM interface by reducing the conductivity mismatch between \nthe MI and HM layers by introducing atomically thin semiconducting interlayers22–28 and \nenhancing the interfacial spin-mixing conductance .29–31 Hariharan’s group has explored the roles \nof bulk and surface magnetic anisotrop ies in LSSE in different REIG -based MI/HM bilayers12,13, \nwhereas a recent study highlights the influence of damping on SPE and LSSE in a compensated \nferrimagnetic insulator .32 It has also been demonstrated that the LSSE in YIG/Pt bilayers varies \ninversely with intrinsic Gilbert damping of the YIG films, however, the LSSE coefficient does not \nshow any significant correlation with the enhanced damping due to SPE in YIG/Pt bilayers.33 All \nthese studies highlight the important role s of both magnetic anisotropy and Gilbert damping in \nthermally generated magnon propagation in MI/HM bilayers . \n \nBy investigating the YIG thickness dependence of the local LSSE measurements in YIG/Pt, \nGuo et al.34 determined the temperature ( T) dependence of 〈𝜉〉 and found a scaling behavior of \n〈𝜉〉 ∝ 𝑇−1. On the contrary , by employing non -local measurement geometries, Cornelissen et al. \ndemonstrated that the magnon diffusion length for thermally driven magnonic spin currents (𝜆𝑡ℎ𝑚) \nof YIG decreases with decreasing temperature over a broad temperature range .35 Gomez -Perez et \nal. reported similar observation s and demonstrate d that the temperature dependen ce of 𝜆𝑡ℎ𝑚 is \n5 \n independent of the YIG thickness .19 The different trends of the temperature dependent \ncharacteristic critical length scales for thermally generated magnon propagation in YIG observ ed \nby different groups indicate s distinct temperature evolutions of 𝛼 and 𝐾𝑒𝑓𝑓 in the YIG films grown \nby these groups. In other words, different thin film growth conditions and sample dependent \nchanges in the physical properties can give rise to different temperature dependences of both 𝛼 \nand 𝐾𝑒𝑓𝑓 and hence 〈𝜉〉. For the development of a complete physical picture of LSSE in these \nREIGs over a wide temperature range, it is of utmost importance to comprehend the respective \nroles of both 𝛼 and 𝐾𝑒𝑓𝑓 simultaneously in determining the temperature evolution of 〈𝜉〉, which \nremains largely unexplored. \n \nAlthough YIG is considered as a benchmark system for LSSE,11,34 there is only a limited \nnumber of studies that explore temperature dependent LSSE in other iron garnets .12,32,36,37 For \nexample, Gd3Fe5O12 (GdIG) which is a ferrimagnetic insulator with magnetic compensation \ntemperature (𝑇𝐶𝑜𝑚𝑝 ) close to room temperature, shows a sign -inversion in the LSSE voltage12 as \nwell as in the spin -Hall anomalous Hall effect38 around its magnetic compensation. However, the \nGilbert damping in GdIG diverges over a broad temperature range around its 𝑇𝐶𝑜𝑚𝑝 which makes \nit difficult to probe the temperature evolution of 𝛼 and its contribution towards 〈𝜉〉 over a wide \ntemperature range around the 𝑇𝐶𝑜𝑚𝑝 .32 Apart from YIG and GdIG , there has been a renaissance of \nresearch interest in another member of the REIG family: Tm 3Fe5O12 (TmIG) due to its wide -\nranging extraordinary magnetic properties6 e.g., strain -tunable perpendicular magnetic anisotropy \n(PMA),39 chiral and topological spin textures,40 and interfacial Dzyaloshinskii -Moriya \ninteraction40,41 combined with low coercivity6 which make this system a promising candidate for \nnumerous efficient spintronic applications, such as spin -orbit torque induced magnetization \n6 \n switching,8,42 current -induced domain -wall motion,43 and spin Hall –topological Hall effect s44,45. \nRecently, the LSSE has been investigated in TmIG/Pt bilayers with PMA at room temperature, \nand shown to exhibit high interfacial spin transparency and spin -to-charge conversion efficiency \nat the TmIG/Pt interface46. TmIG has a higher Gilbert damping parameter (≈10−2)6 compared to \nYIG, and unlike GdIG, TmIG does not exhibit any magnetic compensation in the temperature \nrange between 1.5 and 300 K47,48, which allows us to probe the relative contribution of 𝛼 towards \nthe temperature evolution of 〈𝜉〉 and hence the LSSE over a broad temperature range close to the \nroom temperature . However, the temperature evolution of LSSE and hence 〈𝜉〉 as well as their \nrelation ship with 𝛼 and 𝐾𝑒𝑓𝑓 in TmIG/Pt bilayers are yet to be explored , which would be of critical \nimportance for REIG -based efficient magnonic device applications . Here , we have performed a \ncomprehensive investigation of the temperature -dependent LSSE, radio frequency (RF) transverse \nsusceptibility (TS), and broadband ferromagnetic resonance (FMR) of TmIG /Pt bilayers grown on \ndifferent substrates . From the TmIG thickness dependence of the LSSE voltage, we determined \nthe temperature dependence of 〈𝜉〉 and highlighted its correlation with the temperature -dependent \neffective magnetic anisotropy field, 𝐻𝐾𝑒𝑓𝑓 (∝𝐾𝑒𝑓𝑓) and 𝛼 in TmIG/Pt bilayers. \n \n2. RESULTS AND DISCUSSION \n2. 1. Structural Characterization \nSingle -crystalline TmIG films with different thicknesses were grown on (111) -oriented \nGd3Sc2Ga3O12 (GSGG) and Gd 3Ga5O12 (GGG) substrates by pulsed laser deposition ( see \nMethods ). The high crystalline quality of the TmIG films was confirmed by X-ray diffraction \n(XRD). Figure 1 (a) shows the 𝜃−2𝜃 X-ray diffractograms of the GSGG/TmIG( 𝑡) films with \ndifferent TmIG film thickness 𝑡 (t = 236, 150, 89, 73, 46 and 28 nm) . \n7 \n \n \nFigure 1 . Structural and Morphological characterization. (a) 𝜃−2𝜃 X-ray diffractogram of \nthe GSGG/TmIG( 𝑡) films with different film thickness 𝑡 (t = 236, 150, 89, 73, 46 and 28 nm). The \nreciprocal space maps recorded in the vicinity of the (642) reflection for (b) GSGG/TmIG( 30 nm) \nand (c) GSGG/TmIG(205 nm) films . For the thinner film (30 nm), the TmIG film peak matches \nthe IP lattice constant of the GSGG substrate, whereas for the thicker film (205nm), the TmIG film \nis largely relaxed. \n \n The substrate choice and the TmIG film thickness influence the strain state of the film . \nFigs. 1 (b) and ( c) show the reciprocal space maps in the vicinity of the (642) reflection for the \nGSGG/TmIG ( 30 nm) and GSGG/TmIG (205 nm) films, respectively . For the thinner film (30 \nnm), the TmIG film 𝑞𝑥 matches the in -plane (IP) lattice spacing of the GSGG substrate indicating \ncoherent growth , and the out -of-plane (OOP) lattice spacing is smaller than that of the substrate \n(higher 𝑞𝑍), consistent with the smaller unit cell volume for TmIG compared to GSGG. However, \nthe thicker film (205 nm) is relaxed in plane with smaller IP and OOP lattice spacing than that of \nthe substrate, and its peak position is close to that of bulk TmIG. The 𝜃−2𝜃 scans show a decrease \nin the OOP spacing (increase in 2𝜃) for thinner films . These trends are consistent with the TmIG \n \n8 \n initially growing with an IP lattice match to the substrate and hence a tensile IP strain (and a \nmagnetoelastic anisotropy favoring PMA), but the strain relaxes as the film thickness increases. \nThe thickest films, which are strain -relaxed, have a slightly higher OOP lattice spacing compared \nto bul k according to Fig. 1 (a) which suggests the presence of oxygen vacancies or Tm:Fe ratio \nexceeding 0.6, which can occur in thin films and raise the unit cell volume. All the films show a \nsmooth surface morphology with a low root -mean -square roughness below 0.5nm, as visible in \natomic force microscopy (AFM) images for the GSGG/TmIG( 46nm), GGG/TmIG(44nm) and \nsGGG/TmIG(75nm) films shown in the Supplementary Fig ure 1. \n \n A cross -section of an about 220 nm thick TmIG film on GSGG substrate, covered with \na 5 nm Pt layer, was analyzed by scanning transmission electron microscopy (STEM). Fig. 2 (a) \nshows a low magnification STEM image of the whole layer stack. An annular detector with a small \ncollector angle (24 -48 mrad) was used to highlight strain (Bragg) contrast over mass (Z) contrast \n49. The TmIG film shows columnar features attributed to strain contrast. An atomically resolved \nSTEM image at the TmIG film -Pt interface ( Fig. 2 (b)) reveals a single crystalline TmIG film under \nthe polycrystalline Pt layer, with the bright spots indicating columns of Tm and Fe. The STEM \nimage of an area within the TmIG film close to the Pt interface shows the presence of a planar \ndefect in which s elected lattice planes are highlighted by colored lines in Fig. 2 (c). Such planar \ndefects could be associated with partial dislocations or atomic level disorder, which are common \nin REIGs.50,51 \n \n \n9 \n \n \nFigure 2. Cross -sectional scanning transmission electron microscopy (STEM) analysis of the \nGSGG/TmIG(220 nm)/Pt(5nm) film. (a) TEM image of the layer stack recorded by an annular \ndetector with a small collector angle (24 -48 mrad), highlighting strain (Bragg) contrast over mass \n(Z) contrast, (b) shows an atomic -resolution STEM image of the TmIG -Pt interface with [110] \nzone axis , while (c) shows an area within the TmIG film. The colored lines highlight a planar \ndefect . (d) electron energy loss spectroscopy ( EELS) scan at the Fe L3 and L2 edges. The measured \nenergy loss spectra are displayed as data points, exempl ified for positions close to the garnet -\nsubstrate and garnet -Pt interfaces, with the fitted functions presented as colored lines. (e) The \nthickness dependent Fe L3 peak position and FWHM is extracted. \n \n \n \n \n \n10 \n 2. 2. Correlation between Thermo -Spin Transport and Magnetism \nFig. 3(a) shows the schematic illustration of our LSSE measurement configuration . Simultaneous \napplication of a vertical (+ z-axis) T-gradient ( 𝛁𝑻⃗⃗⃗⃗⃗ ) and an in-plane (x-axis) DC magnetic field \n(𝝁𝟎𝑯⃗⃗⃗⃗⃗⃗⃗⃗ ) across the TmIG fil m causes diffusion of thermally -excited magnons and develops a spatial \ngradient of magnon accumulation along the direction of 𝛁𝑻⃗⃗⃗⃗⃗ .52 The accumulated magnons close to \nthe TmIG/Pt interface transfer spin angular momenta to the electrons of the adjacent Pt layer52. \nThe injected spin current density is, 𝑱𝑺⃗⃗⃗ ∝−𝑆𝐿𝑆𝑆𝐸𝛁𝑻⃗⃗⃗⃗⃗ , where 𝑆𝐿𝑆𝑆𝐸 is the LSSE coefficient52,53. The \nspin current injected into the Pt layer along the z-direction is converted into a charge current , 𝑱𝑪⃗⃗⃗ =\n (2𝑒\nℏ)𝜃𝑆𝐻𝑃𝑡(𝑱𝑺⃗⃗⃗ × 𝝈𝑺⃗⃗⃗⃗⃗ ) along the y-direction via the inverse spin Hall effect (ISHE), where e, ℏ, 𝜃𝑆𝐻𝑃𝑡, \nand 𝝈𝑺⃗⃗⃗⃗⃗ are the electron ic charge, the reduced Planck’s constant , the spin Hall angle of Pt, and the \nspin-polarization vector, respectively . The corresponding LSSE voltage is52,54,55 \n 𝑉𝐿𝑆𝑆𝐸= 𝑅𝑦𝐿𝑦𝐷𝑃𝑡(2𝑒\nℏ)𝜃𝑆𝐻𝑃𝑡| 𝐽𝑆|tanh(𝑡𝑃𝑡\n2𝐷𝑃𝑡), (1) \nwhere, 𝑅𝑦,𝐿𝑦,𝐷𝑃𝑡,and 𝑡𝑃𝑡 represent the electrical resistance between the contact -leads, t he \ndistance between the contact -leads, the spin diffusion length of Pt, and the Pt layer thickness , \nrespectively . \n \nFig. 3(b) shows the magnetic field (H) dependen t ISHE voltage, 𝑉𝐼𝑆𝐻𝐸(𝐻) for \nGSGG/TmIG(236 nm)/Pt (5 nm) for different values of the temperature difference between the hot \n(𝑇ℎ𝑜𝑡) and cold ( 𝑇𝑐𝑜𝑙𝑑) blocks, ∆𝑇=(𝑇ℎ𝑜𝑡−𝑇𝑐𝑜𝑙𝑑), at a fixed average sample temperature 𝑇=\n 𝑇ℎ𝑜𝑡+𝑇𝑐𝑜𝑙𝑑\n2 = 295 K. For all Δ𝑇, 𝑉𝐼𝑆𝐻𝐸(𝐻) exhibit s a nearly square -shaped hysteresis loop. The inset \nof Fig. 3(b) plots the ∆𝑇-dependence of the background -corrected LSSE voltage, 𝑉𝐿𝑆𝑆𝐸(Δ𝑇)= \n11 \n [𝑉𝐼𝑆𝐻𝐸(+𝜇0𝐻𝑠𝑎𝑡,Δ𝑇)−𝑉𝐼𝑆𝐻𝐸(−𝜇0𝐻𝑠𝑎𝑡,Δ𝑇)\n2], where 𝜇0𝐻𝑠𝑎𝑡 is the saturation field . Clearly, 𝑉𝐿𝑆𝑆𝐸 increases \nlinearly with ∆𝑇 as expected from Eqn. 1 .12 \n \nFigure 3. Magnetism and longitudinal spin Seebeck effect (LSSE) in \nGSGG/TmIG(236nm)/Pt(5nm) film. (a) Schematic illustration of the experimental configuration \nfor LSSE measurements. A temperature gradient ( 𝛁𝑻⃗⃗⃗⃗⃗ ) is applied along the + z axis and an in -plane \n(IP) dc magnetic field ( 𝝁𝟎𝑯⃗⃗⃗⃗⃗⃗⃗⃗ ) is applied along the + x axis. The inverse spin Hall effect (ISHE) \ninduced voltage ( 𝑉𝐼𝑆𝐻𝐸) is measured along the y-axis. (b) 𝑉𝐼𝑆𝐻𝐸(𝐻) loops for different values of \nthe temperature difference ∆𝑇 at a fixed average sample temperature 𝑇 = 295 K. The inset shows \na linear ∆𝑇-dependence of the background -corrected LSSE voltage. (c) 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis loops \nmeasured at selected temperatures in the range 120 K ≤ T ≤ 295 K for Δ𝑇 = +10 K. (d) The IP \nM(H) hysteresis loops at selected temperatures. \n \n \n12 \n Fig. 3 (c) shows the 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis loops for GSGG/TmIG(236 nm)/Pt(5 nm) \nmeasured at selected temperatures for Δ𝑇= +10 K. Clearly, |𝑉𝐼𝑆𝐻𝐸(𝜇0𝐻𝑠𝑎𝑡)| significantly \ndecreases, and the hysteresis loop broadens at low temperatures, especially below 200 K. To \ncorrelate thermo -spin transport with the bulk magnetic properties, in Fig. 3 (d), we show the \nmagnetic field dependence of magnetization, 𝑀(𝐻) at selected temperatures for GSGG/TmIG(236 \nnm)/Pt(5 nm) measured while scanning an in -plane (IP) magnetic field. It is evident that with \nlowering the temperature, the saturation magnetization (𝑀𝑆) decreases and t he coercivity ( 𝐻𝐶) \nincreases with a corresponding increase in the magnetic anisotropy, especially below 200 K. This \nobservation is also in agreement with the T-dependent magnetic force microscopy (MFM) results \nshown in Supplementary Figure 2 , which clearly reveals that the root mean square (RMS) value \nof the phase shift, Δ𝜙𝑅𝑀𝑆 decreases significantly between 300 and 150 K indicating changes in the \nmagnetic domain structure at low -T. \n \nThe decrease in 𝑀𝑆 at low -T is well-known in TmIG48,56 and is a result of the increasing \nmoment of the Tm3+ ion at low -T, which competes with the net moment of the Fe3+ ions ( i.e., the \ndodecahedral Tm3+ moment opposes the net moment of the tetrahedral and octahedral Fe3+ \nmoments). Based on the molecular -field-coefficient theory developed by Dionne57, we have \nperformed molecular -field simulations58,59 to determine 𝑀𝑆(T) for TmIG ( see Supplementary \nFigure 3(n)) which is consistent with our experimental observation of the decrease in 𝑀𝑆 at low -\nT. It is apparent from Figs. 3 (c) and (d) that the temperature evolution of 𝑉𝐼𝑆𝐻𝐸 signal follows that \nof 𝑀𝑆. To further explore the correlation between 𝑉𝐼𝑆𝐻𝐸 and 𝑀𝑆, magnetometry and LSSE \nmeasurements were repeated on the GSGG/TmIG( t)/Pt(5 nm) sample series with different TmIG \nfilm thicknesses (28 nm≤𝑡≤236 nm). Films with 46 nm≤𝑡≤236 nm possess IP easy -axes \n13 \n while the 28 nm film has an OOP easy -axis of magnetization , which was confirmed via IP -\nmagnetometry and OOP p-MOKE measurements (see Supplementary Figure 3(e)). The total \nmagnetic anisotropy of a (111) -oriented TmIG fi lm, neglecting growth and interfacial anisotropies, \nhas contributions from shape anisotropy ( 𝐾𝑠ℎ𝑎𝑝𝑒), cubic magnetocrystalline anisotropy ( 𝐾𝑚𝑐), and \nmagnetoelastic anisotropy ( 𝐾𝑚𝑒)47,49,60 i.e., 𝐾𝑒𝑓𝑓=𝐾𝑠ℎ𝑎𝑝𝑒+𝐾𝑚𝑐 + 𝐾𝑚𝑒=−1\n2 𝜇0𝑀𝑆2−𝐾1\n12−\n9\n4𝜆111𝑐44(𝜋\n2−𝛽), where K1 is the magnetocrystalline anisotropy coefficient, 𝜆111 is the \nmagnetostriction along the [111] direction, 𝑐44 is the shear modulus and 𝛽 is the cor ner angle of \nthe rhombohedrally -distorted unit cell. For a negative magnetostriction ( 𝜆111 = −5.2×10−6 for \nbulk TmIG47), the tensile IP strain, which results from the difference in lattice parameters \n(𝑎𝐺𝑆𝐺𝐺=12.57 Å and 𝑎𝑇𝑚𝐼𝐺=12.32 Å) promotes PMA ( 𝐾𝑒𝑓𝑓>0).49,60,61 PMA is expected for \nfully -strained films (28 nm), but strain -relaxation in thicker films reduces the magnetoelastic \ncontribution , and the easy -axis reorients to IP direction60. \n \nFigs. 4 (a) and (b) depict the 𝑉𝐼𝑆𝐻𝐸(𝐻) loop on the left y-scale and corresponding 𝑀(𝐻) \nloop on the right y-scale at 295 K for the thicknesses: 𝑡=236 and 28 nm,respectively . The \n𝑀(𝐻) and 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis -loops for all other thicknesses are shown in the Supplementary \nFigures 3 and 4. Clearly, the 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis -loops for all the thicknesses mimic the \ncorresponding 𝑀(𝐻) loops. Note that, unlike YIG -slab, there is no surface magnetic anisotropy \ninduced anomalous low field feature in the 𝑉𝐼𝑆𝐻𝐸(𝐻) loop for any of our TmIG thin films. This is \npossibly because the thickness of the TmIG films is smaller than their average magnetic domain \nsize62. This is why the YIG thin films also do not show any low field anomalous feature in the \n𝑉𝐼𝑆𝐻𝐸(𝐻) loops52. Additionally, the 𝑉𝐼𝑆𝐻𝐸(𝐻) loop for our TmIG film with PMA ( the 28 nm film) \nat 295 K is quite similar to that of a TmIG thin film with PMA at room temperature reported in the \n14 \n literature46. In Figs. 4 (c) and (d), we demonstrate the T-dependence of the background -corrected \nLSSE voltage, 𝑉𝐿𝑆𝑆𝐸(𝑇)=𝑉𝐼𝑆𝐻𝐸(𝑇,+𝜇0𝐻𝑠𝑎𝑡)−𝑉𝐼𝑆𝐻𝐸(𝑇,−𝜇0𝐻𝑠𝑎𝑡)\n2 for Δ𝑇= +10 K on the left y-scale and \ncorresponding 𝑀𝑆(𝑇) on the right y-scale for GSGG/TmIG( 236 nm )/Pt(5 nm) and \nGSGG/TmIG( 28 nm )/Pt(5 nm) , respectively. Interestingly, 𝑉𝐿𝑆𝑆𝐸(𝑇) and 𝑀𝑆(𝑇) for both films \ndrop remarkably below the T-window of 180 -200 K. We observed a similar trend in 𝑉𝐿𝑆𝑆𝐸(𝑇) and \n𝑀𝑆(𝑇) for all GSGG/TmIG( t)/Pt(5 nm) films with other thicknesses (see Supplementary Figures \n3 and 4 ). These results indicate that this behavior is intrinsic to TmIG. \n \nFigure 4. Longitudinal spin Seebeck effect in GSGG/TmIG( t)/Pt(5 nm) films. The 𝑉𝐼𝑆𝐻𝐸(𝐻) \nhysteresis loops on the left y-scale and the IP 𝑀(𝐻) loops on the right y-scale at T = 295 K for \nGSGG/TmIG( t)/Pt films for t = (a) 236 nm, and (b) 28 nm . The temperature dependence of the \nbackground -corrected LSSE voltage, 𝑉𝐿𝑆𝑆𝐸(𝑇) on the left y-scale and temperature dependence of \nsaturation magnetization, 𝑀𝑆(𝑇) on the right y-scale for the GSGG/TmIG( t)/Pt(5 nm) films for t = \n(c) 236 nm and (d) 28 nm, for Δ𝑇 = +10 K . \n \n15 \n Note that the yellow and grey background colors in all the graphs throughout the \nmanuscript are used to highlight significant changes in physical parameters between high (yellow) \nand low (grey) temperature regions. Additionallly, we have used the sky blue background color in \nsome of the specific graphs (especially temperature dependence of 𝑉𝐿𝑆𝑆𝐸(𝑇) and 𝑀𝑆(𝑇)) to \nindicate considerable changes in the corresponding physical parameters occurring around the \nnarrow temperature window: 180 K ≤𝑇 ≤200 K. However, we have used a gradual transition \nfrom yellow to grey background in rest of the graphs where the changes in the physical parameters \nare less significant in the temperature window: 180 K ≤𝑇 ≤200 K. \n \nNext, we discuss the additional voltage contribution s due to the magnetic proximity effect \n(MPE) -induced anomalous Nernst effect (ANE) as well as MPE –induced LSSE in the Pt layer. \nThe MPE leads to a magnetic moment in a few atomic layers of Pt close to the TmIG/Pt \ninterface.63,64 In the presence of a vertical temperature gradient, a tra nsverse voltage is generated \nin the proximitized Pt layer due to ANE which adds to the LSSE voltage. Furthermore, due to the \ntemperature gradient, spin currents are generated inside the magnetized Pt layer, which induces an \nadditional IP charge current at the proximitized Pt/nonmagnetic Pt interface via the ISHE and \ntherefore contributes to the LSSE signa l.65 In an earlier study, Bougiatioti et al.,63 showed that the \nMPE -induced ANE in the proximitized Pt layer is only significant for a conducting FM/Pt bilayer \nbut negligible for semiconducting FM/Pt bilayers and becomes zero for insulating FM/Pt bilayers. \nSince TmIG is insulating, the contribution of the MPE -induced ANE in the proximitized Pt layer \ntowards the total LSSE signal can be neglected throughout the measured temperature range66. \nFurthermore , since the LSSE voltage decreases with decreasing thickness of the magnetic layer ,16 \nand the thickness of the proximitized Pt layer is very small, the MPE -induced LSSE contribution \n16 \n due to the proximitized Pt layer can also be neglected66. Therefore, the total voltage measured \nacross the TmIG/Pt bilayers is considered to be solely contributed by the intrinsic LSSE of the \nTmIG films. \n \n2. 3. Analysis of the Thickness Dependent Longitudinal Spin Seebeck Effect \nTo ascertain the origin of the decrease in 𝑉𝐿𝑆𝑆𝐸 below 180 -200 K in our TmIG films, it is \nessential to determine the temperature evolution of 〈𝜉〉 which signifies the critical length -scale for \nthe thermally -generated magnons of a magnetic thin film16,18,34. For an effective determination of \nthe temperature dependence of 〈𝜉〉, the contributions of the the thermal resistances of the substrate \nand the grease layers as well as the interfacial thermal resistances need to be considered.67 To \nquantify the temperature evolution of 〈𝜉〉 for our TmIG/Pt bilayer films, we have employed a \nmodel proposed by Jimenez -Cavero et al.,68 according to which t he total temperature difference \n(Δ𝑇) across the GSGG/TmIG/Pt heterostructure can be expressed as a linear combination of \ntemperature drops in the Pt layer, at the TmIG/Pt interface, in the TmIG layer, at the GSGG/TmIG \ninterface and across the GSGG substrate as well as in the N -grease layers (thickness ≈ 1 m) on \nboth sides of the GSGG/TmIG/Pt heterostructures as ,68 ∆𝑇= ∆𝑇𝑃𝑡+∆𝑇Pt\nTmIG+∆𝑇TmIG+\n ∆𝑇TmIG\nGSGG+∆𝑇GSGG+2.∆𝑇N−Grease (see Fig. 5 (a)). Assuming negligible drops in ∆𝑇 in the Pt layer \nand at the GSGG/TmIG and Pt/TmIG interface ,68,69 the total temperature difference can be \napproximately written as, ∆𝑇= ∆𝑇Pt\nTmIG+∆𝑇TmIG+∆𝑇GSGG+2.∆𝑇𝑁−𝐺𝑟𝑒𝑎𝑠𝑒 . Considering these \ncontributions, the temperature drops in the TmIG layer and at the TmIG/Pt interface can be written \nas,68,69 ∆𝑇TmIG = 𝛥𝑇\n [1+𝜅𝑇𝑚𝐼𝐺\n𝑡𝑇𝑚𝐼𝐺(2𝑡𝑁−𝐺𝑟𝑒𝑎𝑠𝑒\n𝜅𝑁−𝐺𝑟𝑒𝑎𝑠𝑒 + 𝑡𝐺𝑆𝐺𝐺\n𝜅𝐺𝑆𝐺𝐺)] and ∆𝑇Pt\nTmIG= [(𝜅𝐺𝑆𝐺𝐺𝜅𝑇𝑚𝐼𝐺)𝑅𝑖𝑛𝑡\n(𝜅𝑇𝑚𝐼𝐺𝑡𝐺𝑆𝐺𝐺+𝜅𝐺𝑆𝐺𝐺𝑡𝑇𝑚𝐼𝐺)]𝛥𝑇, \nrespectively. The bulk (𝑉𝐿𝑆𝑆𝐸𝑏)and interfacial (𝑉𝐿𝑆𝑆𝐸𝑖) contributions to the LSSE voltage can then \n17 \n be expressed as, 𝑉𝐿𝑆𝑆𝐸𝑏=𝑆𝐿𝑆𝑆𝐸𝑏.∆𝑇TmIG.𝐿𝑦=\n[(𝐴\n𝑡𝑇𝑚𝐼𝐺){cosh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)−1\nsinh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)}]{𝛥𝑇\n[1+𝜅𝑇𝑚𝐼𝐺\n𝑡𝑇𝑚𝐼𝐺(2𝑡𝑁−𝐺𝑟𝑒𝑎𝑠𝑒\n𝜅𝑁−𝐺𝑟𝑒𝑎𝑠𝑒 + 𝑡𝐺𝑆𝐺𝐺\n𝜅𝐺𝑆𝐺𝐺)]}𝐿𝑦 and 𝑉𝐿𝑆𝑆𝐸𝑖=𝑆𝐿𝑆𝑆𝐸𝑖.∆𝑇Pt\nTmIG.𝐿𝑦=\n𝑆𝐿𝑆𝑆𝐸𝑖.[(𝜅𝐺𝑆𝐺𝐺𝜅𝑇𝑚𝐼𝐺)𝑅𝑖𝑛𝑡𝛥𝑇\n(𝜅𝑇𝑚𝐼𝐺𝑡𝐺𝑆𝐺𝐺+𝜅𝐺𝑆𝐺𝐺𝑡𝑇𝑚𝐼𝐺)]𝐿𝑦, respectively. Here, 𝑆𝐿𝑆𝑆𝐸𝑏=[(𝐴\n𝑡𝑇𝑚𝐼𝐺){cosh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)−1\nsinh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)}] and 𝑆𝑖𝑛𝑡 \ndenote the bulk and interfacial LSSE coefficient s for TmIG and TmIG/Pt interface , respectively, \n𝑡𝑇𝑚𝐼𝐺(𝑡𝐺𝑆𝐺𝐺) is the thickness of TmIG film (GSGG substrate), 𝜅𝑇𝑚𝐼𝐺 and 𝜅𝐺𝑆𝐺𝐺 are the thermal \nconductivity of TmIG and GSGG respectively, 𝑡N−Grease and 𝜅N−Grease are the thickness and \nthermal conductivity of the N -grease layers, 𝑅𝑖𝑛𝑡 is the interfacial thermal -resistance at the \nTmIG/Pt interface and 𝐴 is a constant .68. The approximate values of 𝜅N−Grease , 𝜅𝑇𝑚𝐼𝐺 and 𝜅𝐺𝑆𝐺𝐺 \nat different temperatures are obtained from the literature70–75. Note that , we have ignored the \ninterfacial thermal resistances between the N -grease and the hot/cold plates as well as between the \nsample and N -grease layers .76 Threfore , the total LSSE voltage across GSGG/TmIG/Pt can be \nexpressed as,68 \n𝑉𝐿𝑆𝑆𝐸(𝑡𝑇𝑚𝐼𝐺)= 𝑉𝐿𝑆𝑆𝐸𝑖(𝑡𝑇𝑚𝐼𝐺)+𝑉𝐿𝑆𝑆𝐸𝑏(𝑡𝑇𝑚𝐼𝐺)= [𝑆𝑖𝑛𝑡{(𝜅𝐺𝑆𝐺𝐺𝜅𝑇𝑚𝐼𝐺)𝑅𝑖𝑛𝑡\n(𝜅𝑇𝑚𝐼𝐺𝑡𝐺𝑆𝐺𝐺+𝜅𝐺𝑆𝐺𝐺𝑡𝑇𝑚𝐼𝐺)}𝐿𝑦𝛥𝑇+\n[(𝐴\n𝑡𝑇𝑚𝐼𝐺){cosh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)−1\nsinh(𝑡𝑇𝑚𝐼𝐺\n〈𝜉〉)}]{𝛥𝑇\n[1+𝜅𝑇𝑚𝐼𝐺\n𝑡𝑇𝑚𝐼𝐺(2𝑡𝑁−𝐺𝑟𝑒𝑎𝑠𝑒\n𝜅𝑁−𝐺𝑟𝑒𝑎𝑠𝑒 + 𝑡𝐺𝑆𝐺𝐺\n𝜅𝐺𝑆𝐺𝐺)]}𝐿𝑦] (2) \n \nIn Fig. 5 (b), we demonstrate the 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis loops at T = 295 K for the \nGSGG/TmIG/Pt films with different 𝑡𝑇𝑚𝐼𝐺 in the range 28 nm≤𝑡≤236 nm. Clearly, \n|𝑉𝐼𝑆𝐻𝐸(𝜇0𝐻𝑠𝑎𝑡)| decreases significantly with decreasing 𝑡𝑇𝑚𝐼𝐺 . Therefore, we fitted the thickness \ndependent LSSE voltage at different temperatures with Eqn. 2 to evaluate the temperature \ndependence of 〈𝜉〉 for our GSGG/TmIG/Pt films. It has recently been shown77 that the 𝑀𝑆 also \n18 \n needs be considered to evaluate 〈𝜉〉 from the LSSE voltage by normalizing the LSSE voltage by \n𝑀𝑆. In Fig. 5 (c), we show the thickness -dependence of the background -corrected modified LSSE \nvoltage,𝑉𝐿𝑆𝑆𝐸(𝑡𝑇𝑚𝐼𝐺)\n∆𝑇.𝑀𝑆, at selected temperatures fitted to Eqn. 2 . From the fits, we obtained 〈𝜉〉 = 62 \n± 5 nm for the TmIG film at 295 K, which is smaller than that of YIG thin films grown by PLD \n(90–140 nm)16, but higher than that for GdIG thin films (45±8 nm)12. \n \nFigure 5. Thickness D epende nt LSSE and Magnon Propagation L ength in \nGSGG/TmIG( t)/Pt(5 nm) films. (a) Schematic illustration of heat flow across the \nGSGG/TmIG( t)/Pt(5 nm) films. (b) The 𝑉𝐼𝑆𝐻𝐸(𝐻) hysteresis loops for GSGG/TmIG( t)/Pt films \nwith different thicknesses at T = 295 K for Δ𝑇 = +10 K . (c) The thickness dependence of the \nnormalized background corrected LSSE voltage, 𝑉𝐿𝑆𝑆𝐸(𝑡)Δ𝑇.𝑀𝑆⁄ at three selected temperatures \nT = 295, 200 , and 140 K fitted with Eqn. (2) . (d) The temperature dependence of the magnon \npropagation length, 〈𝜉〉 obtained from the fits. \n \n19 \n Fig. 5 (d) demonstrates the T-dependence of 〈𝜉〉 obtained from the fit of 𝑉𝐿𝑆𝑆𝐸(𝑡𝑇𝑚𝐼𝐺)\n∆𝑇.𝑀𝑆 for \nGSGG/TmIG( t)/Pt(5nm) films . Interestingly, 〈𝜉〉 decreases gradually with decreasing temperature \nand shows a comparatively faster decrease at low temperatures, especially below 200 K. Our \nobservation is strikingly different than that reported by Guo et al.34 for YIG/Pt bilayers. From the \nYIG thickness dependence of the local LSSE measurements in YIG/Pt, they determined the \ntemperature dependence of 〈𝜉〉 and found a scaling behavior of 〈𝜉〉 ∝ 𝑇−1.34 However, by \nemploying nonlocal measurement geometries, Cornelissen et al., demonstrated that 〈𝜉〉 (and hence, \nthe magnon diffusion length) for YIG/Pt decreases with decreasing temperature over a broad \ntemperature range35, similar to what we observed in our TmIG/Pt bilayers. Gomez -Perez et al.,19 \nalso observed similar behavior of the magnon diffusion length in YIG/Pt. However, none of these \nstudies indicated significant change in 〈𝜉〉 at low temperatures . Therefore, the observed \ntemperature evolution of 〈𝜉〉 presented in this study is intrinsic to TmIG. To rule out possible \neffects of strain o n 𝑉𝐿𝑆𝑆𝐸(𝑇), we performed LSSE measurements on TmIG films grown on \ndifferent substrates (see Supplementary Figures 5 and 6). It is evident that 𝑀𝑆(𝑇) and 𝑉𝐿𝑆𝑆𝐸(𝑇) \nfor the Gd3Ga5O12(GGG) /TmIG(44 nm)/Pt (5 nm) and \n(Gd 2.6Ca0.4)(Ga 4.1Mg 0.25Zr0.65)O12(sGGG )/TmIG(40 nm)/Pt (5 nm) films (see Supplementary \nFigure 7) exhibit the same trend as GSGG/TmIG(46 nm)/Pt(5 nm). More specifically , both \n𝑉𝐿𝑆𝑆𝐸(𝑇) and 𝑀𝑆(𝑇) drop be low 180 -200 K for all the TmIG films independent of substrate choice. \n \nTo interpret the decrease in 〈𝜉〉 at lo w temperatures , we recall that 〈𝜉〉 of a magnetic \nmaterial with lattice constant 𝑎0 (considering simple cubic structure) is related to the Gilbert \ndamping parameter ( 𝛼), the effective magnetic anisotropy constant ( 𝐾𝑒𝑓𝑓), and the strength of the \n20 \n Heisenberg exchange interaction between nearest neighbors ( 𝐽𝑒𝑥) through the relation17,18 〈𝜉〉=\n 𝑎0\n2𝛼.√𝐽𝑒𝑥\n2𝐾𝑒𝑓𝑓. As discussed before, 𝐾𝑒𝑓𝑓=𝐾𝑚𝑒−1\n2 𝜇0𝑀𝑆2−𝐾1\n12. Therefore, we can express 〈𝜉〉 as, \n〈𝜉〉= 𝑎0\n2𝛼.√𝐽𝑒𝑥\n2(𝐾𝑚𝑒−𝐾1\n12−1\n2 𝜇0𝑀𝑆2) (3) \nEqn. 3 indicates that (i) 〈𝜉〉∝ (1\n𝛼), and (ii) a decrease in 𝑀𝑆 also suppresses 〈𝜉〉. Since the \neffective anisotropy field, 𝐻𝐾𝑒𝑓𝑓 ∝𝐾𝑒𝑓𝑓, Eqn. 3 can be alternatively written as 〈𝜉〉=𝑎0\n2𝛼.√𝐽𝑒𝑥\n2𝐾𝑒𝑓𝑓 ∝\n1\n𝛼.(𝐻𝐾𝑒𝑓𝑓)1/2 , which indicates that 〈𝜉〉 is inver sely proportional to the square -root of 𝐻𝐾𝑒𝑓𝑓. This \nimplie s that the temperature evolution of 〈𝜉〉 is intrinsically dependent on both the physical \nquantities: 𝛼 and 𝐻𝐾𝑒𝑓𝑓. To determine the roles of 𝛼 and 𝐻𝐾𝑒𝑓𝑓in the temperature evolution of 〈𝜉〉, \nwe have performed radio frequency (RF) transverse susceptibility (TS) and broadband \nferromagnetic resonance (FMR) measurements, respectively on the TmIG films, which have been \ndiscussed in the following sections. \n \n2. 4. Radio Frequency Transverse Susceptibility and Magnetic Anisotropy \nRF TS measurements were performed to determine the temperature evolution of 𝐻𝐾𝑒𝑓𝑓 in \nthe TmIG films. The magnetic field dependence ( 𝐻𝐷𝐶) of TS, 𝜒𝑇(𝐻𝐷𝐶), is known to exhibit \npeaks/cusps at the effective anisotropy fields, ±𝐻𝐾𝑒𝑓𝑓.78,79 The schematic illustration of our TS \nmeasurement configuration is shown in Fig. 6 (a). T he RF magnetic field, HRF is parallel to the film \nsurface and 𝐻𝐷𝐶 points perpendicular to it. All the TS data in this paper are presented as the \nrelative change in 𝜒𝑇(𝐻𝐷𝐶), which we define as ∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶)=𝜒𝑇(𝐻𝐷𝐶)−𝜒𝑇(𝐻𝐷𝐶=𝐻𝐷𝐶𝑠𝑎𝑡)\n𝜒𝑇(𝐻𝐷𝐶𝑠𝑎𝑡), where \n𝜒𝑇(𝐻𝐷𝐶= 𝐻𝐷𝐶𝑠𝑎𝑡) is the value of 𝜒𝑇(𝐻𝐷𝐶) at the saturation field ( 𝐻𝐷𝐶𝑠𝑎𝑡). Bipolar field -scans of \n21 \n ∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶) for the GSGG/TmIG( 236 nm )/Pt film at 295 and 100 K are shown in Fig. 6 (b), which \nclearly indicates an increase in 𝐻𝐾𝑒𝑓𝑓 at low -T. \n \nFigure 6. RF Transverse Susceptibility and Magnetic Anisotropy in GSGG/TmIG( t)/Pt(5 \nnm) films. (a) The schematic illustration of our RF transverse susceptibility measurement. (b) \nComparison of the bipolar field scans ( +𝐻𝐷𝐶𝑚𝑎𝑥→−𝐻𝐷𝐶𝑚𝑎𝑥→+𝐻𝐷𝐶𝑚𝑎𝑥) of transverse susceptibility at \nT = 295 and 100 K for the GSGG/TmIG( 236 nm)/Pt film measured with configuration 𝐻𝐷𝐶⊥\nfilm surface (IP easy axis) . (c) Fitting of our ∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶) data for the GSGG/TmIG( 236nm )/Pt film \nat 295 K with the Eqn. 4 . (d) Temperature dependence of the effective anisotropy field ( 𝐻𝐾𝑒𝑓𝑓) for \nthe GSGG/TmIG(236 nm)/Pt(5 nm) film obtained from the transverse susceptibility (TS) \nmeasurements on the left y-scale and corresponding 𝑉𝐿𝑆𝑆𝐸(𝑇) for the same film on the right y-scale. \n \n \n22 \n For an accurate determination of 𝐻𝐾𝑒𝑓𝑓from the field dependent TS curves, we fitted the \nline shapes for the TS curves with the following expression,79,80 \n∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶)= ∆𝜒𝑆𝑦𝑚(∆𝐻\n2)2\n(𝐻𝐷𝐶−𝐻𝐾𝑒𝑓𝑓)2\n+(∆𝐻\n2)2+∆𝜒𝐴𝑠𝑦𝑚∆𝐻\n2(𝐻𝐷𝐶−𝐻𝐾𝑒𝑓𝑓)\n(𝐻𝐷𝐶 −𝐻𝐾𝑒𝑓𝑓)2\n+(∆𝐻\n2)2+∆𝜒0 (4) \nwhere, ∆𝐻 is the linewidth of the ∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶) spectrum, ∆𝜒𝑆𝑦𝑚 and ∆𝜒𝐴𝑠𝑦𝑚 are the coefficients of \nsymmetric and antisymmetric Lorentzian functions and ∆𝜒0 is the constant offset parameter. Fig. \n6(c) shows the fitting of our ∆𝜒𝑇\n𝜒𝑇(𝐻𝐷𝐶) data for the GSGG/TmIG( 236 nm )/Pt film at 295 K with \nthe Eqn. 4 . As shown on the left y-scale of Fig. 6 (d), 𝐻𝐾𝑒𝑓𝑓(𝑇) increases throughout the measured \ntemperature range but the increase in 𝐻𝐾𝑒𝑓𝑓 is comparatively faster below the temperature range: \n180-200 K, which coincides with the remarkable drop in 𝑉𝐿𝑆𝑆𝐸. Similar behavior was also \nobserved for other film thicknesses (see Supplementary Figure 8 ). \n \nA significant increase in magnetocrystalline anisotropy at low -T has been reported in \nvarious REIGs, which was interpreted in the framework of the single -ion anisotropy model \nconsidering the collective influence of the crystal and exchange fields of the REIG on the energy \nlevels of the individual magnetic ions81. Typically, 𝐾1 increases by ≈ 80 -100% between 300 and \n150 K in most of the REIGs .81 Furthermore, 𝜆111 for TmIG increases from −5.2 ×10−6 at 300 K \nto −17.4 ×10−6 at 150 K which gives rise to enhanced contribution of 𝐾𝑚𝑒 towards 𝐾𝑒𝑓𝑓 in \nTmIG films at low temperatures.82,83 Shumate Jr. et al.,84 observed a rapid increase in 𝐻𝐾𝑒𝑓𝑓 and \ncoercive field at low temperatures in mixed REIGs. An increase in 𝐻𝐾𝑒𝑓𝑓and a corresponding \ndecrease in 𝑉𝐿𝑆𝑆𝐸 below 175 K was also observed in YIG/Pt13, which was attributed to the single -\nion anisotropy of Fe2+ ions85. To gain knowledge on the oxidation state of Fe in our TmIG films, \n23 \n electron energy loss spectroscopy (EELS) was conducted during the cross -sectional TEM study \ndescribed earlier. Fig. 2 (d) shows two EELS spectra, recorded at the Fe L3 and L2 edges, and at \npositions close to the film -substrate and the film -Pt interface. The spectra are fitted following86,87, \nshown by colored lines, using a Gauß ian profile and a combination of a power -law background \nand a double -step function (arctangent) with a fixed step -ratio. Fig. 2 (e) shows the extracted \nthickness -dependent Fe L3 peak position alongside the corresponding FWHM. While an exact \nquantification of the Fe oxidation state distribution using the EELS Fe L3 peak position or L3/L2 \nwhite -line ratio is challenging, the presence of different oxidation states can be indicated \nqualitatively by a shift in the peak position because Fe2+ ions contribute at slightly lower energies \ncompared to Fe3+ ions86–89. However, in our measured spectra, a constant peak position at about \n710.1 eV and a constant FWHM of about 2.3 eV across the whole film thickness is observed. Our \nobservation strongly hints at the presence of only one Fe oxidation state, namely the Fe3+ ion and \nhence, we can rule out the contribution of single ion anisotropy of Fe2+ ions towards the increased \nmagnetic anisotropy. This is also in agreement with recent studies on Tb-rich TbIG thin films58,90 \nwhich reveal very low Fe2+ ion concentrations . Therefore, the increase in 𝐻𝐾𝑒𝑓𝑓 below 200 K in the \nTmIG films may arise from single -ion anisotrop ies of the Tm3+ and Fe3+ ions81 as well as from the \nenhanced contributions of 𝐾1 and 𝐾𝑚𝑒 towards 𝐾𝑒𝑓𝑓 at low temperatures81–83. \n \n2. 5. Magnetization Dynamics and Broadband Ferromagnetic Resonance \nNext, we examine the temperature evolution of 𝛼 and its influence on 〈𝜉〉 through \nbroadband IP FMR measurements. Fig. 7 (a) shows the field -derivative of the microwave (MW) \npower absorption spectra (𝑑𝑃\n𝑑𝐻) as a function of the IP DC magnetic field for a fixed frequency f = \n12 GHz at selected temperatures for the GSGG/ TmIG( 236nm ) film . As temperature decreases, the \n24 \n (𝑑𝑃\n𝑑𝐻) lineshape noticeably broadens and the resonance field 𝐻𝑟𝑒𝑠 shifts to higher field values. The \nlinewidth of the (𝑑𝑃\n𝑑𝐻) lineshape becomes so broad at low temperatures that we were unable to \ndetect the FMR signal below 160 K. We observed the same behavior for the GSGG/TmIG( 236 \nnm)/Pt(5 nm) film, as shown in the Supplementary Figure 9 . Fig. 7 (b) shows the (𝑑𝑃\n𝑑𝐻) lineshapes \nfor the GSGG/TmIG( 236 nm ) film for different frequencies in the range 6 GHz ≤𝑓 ≤20 GHz \nat 295K fitted with a linear combination of symmetric and antisymmetric Lorentzian function \nderivatives as,91 \n 𝑑𝑃\n𝑑𝐻= 𝑃𝑆𝑦𝑚∆𝐻\n2(𝐻𝑑𝑐−𝐻𝑟𝑒𝑠)\n[(𝐻𝑑𝑐−𝐻𝑟𝑒𝑠)2+(∆𝐻\n2)2\n]2+𝑃𝐴𝑠𝑦𝑚(∆𝐻\n2)2\n−(𝐻𝑑𝑐−𝐻𝑟𝑒𝑠)2\n[(𝐻𝑑𝑐−𝐻𝑟𝑒𝑠)2+(∆𝐻\n2)2\n]2+𝑃0 (5) \nwhere, 𝐻𝑟𝑒𝑠 is the resonance field, ∆𝐻 is the linewidth of the 𝑑𝑃\n𝑑𝐻 lineshapes, 𝑃𝑆𝑦𝑚 and \n𝑃𝐴𝑠𝑦𝑚 are the coefficients of the symmetric and antisymmetric Lorentzian derivatives, respectively, \nand 𝑃0 is a constant offset parameter. The fitted curves are shown by solid lines in Fig. 7 (b). Using \nthe values of 𝐻𝑟𝑒𝑠 obtained from the fitting of the 𝑑𝑃\n𝑑𝐻 lineshapes, we fitted the f-𝐻𝑟𝑒𝑠 curves at \ndifferent temperatures using the Kittel equation for magnetic thin films with IP magnetic field,92 \nwhich is expressed as 𝑓= 𝛾𝜇0\n2𝜋√𝐻𝑟𝑒𝑠(𝐻𝑟𝑒𝑠+𝑀𝑒𝑓𝑓), where 𝑀𝑒𝑓𝑓 is the effective magnetization, \n𝛾\n2𝜋= 𝑔𝑒𝑓𝑓 𝜇𝐵\nℏ is the gyromagnetic ratio, 𝜇𝐵 is the Bohr magneton, 𝑔𝑒𝑓𝑓 is the effective Landé g-\nfactor, and ℏ is the reduced Planck’s constant. Fig. 7 (d) demonstrates the fitting of the f-𝐻𝑟𝑒𝑠 \ncurves at T = 295, 200, and 160 K. We found that 𝑔𝑒𝑓𝑓 = 1.642 ± 0.002 at T = 295 K for our \nGSGG/TmIG( 236 nm ) film, which is significantly lower than that of the free electron value ( 𝑔𝑒𝑓𝑓 \n= 2.002), but close to the bulk TmIG value ( 𝑔𝑒𝑓𝑓 = 1.63)93 as well as that for TmIG thin films \n(𝑔𝑒𝑓𝑓 ≈ 1.57)6,94. Furthermore, as shown in Fig. 7( e), 𝑔𝑒𝑓𝑓 for our GSGG/TmIG( 236 nm ) film \n25 \n decreases gradually with decreasing temperature . We observed similar behavior of 𝑔𝑒𝑓𝑓 for the \nGSGG/TmIG( 236 nm )/Pt(5 nm) film, and well as for other TmIG film thicknesses (see \nSupplementary Figures 9 and 10 ). \n \nFigure 7. Broadband Ferromagnetic Resonance. (a) The field derivative of microwave (MW) \npower absorption spectra ( 𝑑𝑃\n𝑑𝐻 line shapes) for the GSGG/TmIG( 236 nm ) film at a fixed \nfrequency ( f = 12 GHz) in the range 160 K ≤ T ≤ 295 K . (b) 𝑑𝑃\n𝑑𝐻 line shapes at different frequencies \nbetween f = 6 - 20 GHz fitted with the linear combination of symmetric and anti -symmetric \nLorentzian function derivatives for the GSGG/TmIG( 236 nm ) film at T = 295 K . (c) Frequency \ndependence of linewidth, ∆𝐻 at different temperatures for the GSGG/TmIG( 236 nm ) film with \nlinear fit. (d) The f-𝐻𝑟𝑒𝑠 curves at T = 295, 200, and 160 K along with Kittel fits. (e) Temperature \ndependence of the Gilbert damping parameter, 𝛼𝑇𝑚𝐼𝐺 , the inhomogeneous broadening, ∆𝐻0 and \nthe effective Landé g-factor for the GSGG/TmIG( 236 nm) film. \n \n \n26 \n Finally, to quantify the temperature dependence of the Gilbert damping parameter (𝛼𝑇𝑚𝐼𝐺 ), \nwe fitted the ∆𝐻-f curves at different temperatures using the expression ,95 ∆𝐻=∆𝐻0+4𝜋𝛼\n𝛾𝜇0𝑓, \nwhere ∆𝐻0 is the frequency -independent contribution to the linewidth, known as the \ninhomogeneous broadening linewidth . From the fits ( see Fig. 7(c)), we obtained 𝛼𝑇𝑚𝐼𝐺 = 0.0103 \n± 0.002 at 295 K for our GSGG/TmIG( 236 nm) film which is close to the previously reported \nvalues of 𝛼 (≈ 0.0132 -0.0146) for TmIG films6,96. Most importantly, 𝛼𝑇𝑚𝐼𝐺 increases gradually \nwith decreasing temperature but shows a comparatively faster increase at low temperatures, \nespecially below ≈ 200 K (Fig. 7(e)). A similar increase in 𝛼 at low-T has also been observed in \nGSGG/TmIG( 236 nm)/Pt(5 nm), GSGG /TmIG( 46 nm)/Pt(5 nm) and GGG/TmIG( 44 nm)/Pt(5 \nnm) films (see Supplementary Figures 9 and 10), indicating that this behav ior is independent of \nTmIG film thickness and substrate choice. In compensated ferrimagnetic insulators, e.g., GdIG, 𝛼 \nincreases drastically close to the magnetic compensation temperature32. However, most of the \nearlier reports indicate that TmIG films do not show magnetic compensation in the temperature \nrange between 1.5 and 300 K.47,48 Since our TmIG films also do not show magnetic compensation \nin the measured temperature range, the increased value of 𝛼𝑇𝑚𝐼𝐺 at low temperatures in our TmIG \nfilms has a different origin . Sizeable increase s in 𝛼 and ∆𝐻 at low temperatures were also observed \nin YIG and different REIGs92,97 –99 including TmIG46, which was primarily attributed to Fe2+ and/or \nRE3+ impurity relaxation mechanisms. However, our EELS study confirms the absence of Fe2+ \nions, and therefore, we can rule out the possibility of Fe2+ impurity relaxation in our TmIG films. \nTherefore, t he increased damping at low temperatures in our TmIG films may be associated with \nenhanced magnon scattering by defects ,100–104 and slowly relaxing Tm3+ ions92,105. It is known that \nthe contribution of slowly relaxing RE impurity ions towards damping is proportional to the orbital \nmoment (L) of the RE3+ ions105–107, suggesting that this mechanism applies to Tm3+ (L = 5). \n27 \n 2. 6. Correlating Magnon Propagation Length with Magnetic Anisotropy and Gilbert \nDamping \nIn the previous sections, we have demonstrated that both 𝐻𝐾𝑒𝑓𝑓and 𝛼𝑇𝑚𝐼𝐺 for our TmIG \nfilms show clear increases at low temperatures, especially below 200 K. It is known that the \nmagnon energy -gap (ℏ𝜔𝑀) is related to 𝐾𝑒𝑓𝑓 through the expression: ℏ𝜔𝑀∝2𝐾𝑒𝑓𝑓17,18. \nTherefore, an increase in 𝐻𝐾𝑒𝑓𝑓 (and hence, 𝐾𝑒𝑓𝑓) below 200 K enhances ℏ𝜔𝑀 giving rise to only \nhigh-frequency magnon propagation with shorter 〈𝜉〉. Since only the subthermal magnons, i.e., the \nlow frequency magnons are primarily responsible for the long -range thermo -spin transport and \ncontributes towards LSSE19–21, the 𝑉𝐿𝑆𝑆𝐸 signal also decreases below 200 K in our TmIG films12,13. \nThis also explains the noticeable decrease in 〈𝜉〉 below 200 K, as the maximum value of the \nfrequency -dependent propagation length is 〈𝜉〉𝑚𝑎𝑥∝1\n√ℏ𝜔𝑀𝑚𝑖𝑛 , where ℏ𝜔𝑀𝑚𝑖𝑛 is the minimum value \nof ℏ𝜔𝑀, and ℏ𝜔𝑀𝑚𝑖𝑛 ∝2𝐾𝑒𝑓𝑓.17 Therefore, according to the expression17,18 〈𝜉〉=𝑎0\n2𝛼.√𝐽𝑒𝑥\n2𝐾𝑒𝑓𝑓 ∝\n1\n𝛼.(𝐻𝐾𝑒𝑓𝑓)1/2 , the observed decrease in 〈𝜉〉 and hence, the 𝑉𝐿𝑆𝑆𝐸 signal at low temperatures, \nespecially below 200 K in our TmIG films has contributions from the temperature evolutions of \nboth 𝐻𝐾𝑒𝑓𝑓and 𝛼𝑇𝑚𝐼𝐺 . The roles of magnetic anisotropy and damping in LSSE in different REIG -\nbased MI/HM bilayers have been explored by different groups12,13,32,33. All these studies indicated \nthat the LSSE signal strength varies inversely with both magnetic anisotropy and damping. In this \nmanuscript, we have not only highlighted the roles of the temperature evolutions of both magnetic \nanisotropy and damping in cont rolling the temperature dependent LSSE effect in TmIG/Pt bilayers, \nbut also attempted to establish possible correlations between 〈𝜉〉, 𝐻𝐾𝑒𝑓𝑓and 𝛼. Since 〈𝜉〉 is intrinsic \nto a magnetic film and hence independent of the thickness of the magnetic film19, it is convenient \n28 \n to directly correlate 〈𝜉〉 with the physical parameters 𝐻𝐾𝑒𝑓𝑓and 𝛼 of individual magnetic films with \ndifferent thicknesses. We display the temperature dependence of 〈𝜉〉 on the left -y scales, and the \ntemperature evolutions of 𝐻𝐾𝑒𝑓𝑓and 𝛼𝑇𝑚𝐼𝐺 are shown on the right y-scales of Figs. 8 (a) and (b), \nrespectively. It is evident that the prominent drop in 〈𝜉〉 below 200 K in the TmIG/Pt bilayers is \nassociated with the noticeable increases in 𝐻𝐾𝑒𝑓𝑓 and 𝛼 that occur within the same temperature \nrange. \n \nFigure 8. Temperature evolution of magnon propagation length and its correlation with \nmagnetic anisotropy and Gilbert damping: (a) and (b) Temperature dependence of 〈𝜉〉 on the \nleft-y scales, and the temperature evolutions of 𝐻𝐾𝑒𝑓𝑓and 𝛼𝑇𝑚𝐼𝐺 are shown on the right y-scales , \nrespectively. 〈𝜉〉 as a function of (c) √𝐻𝐾𝑒𝑓𝑓 and (d) 𝛼𝑇𝑚𝐼𝐺 for the GSGG/TmIG(236 nm) film \nobtained from the temperature evolutions of 〈𝜉〉, 𝐻𝐾𝑒𝑓𝑓 and 𝛼𝑇𝑚𝐼𝐺 . \n \n29 \n For a clearer understanding of the direct correlation between 〈𝜉〉 and 𝐻𝐾𝑒𝑓𝑓 in our TmIG/Pt \nbilayer films , we have plotted 〈𝜉〉 as a function of √𝐻𝐾𝑒𝑓𝑓 for the GSGG/TmIG(236 nm)/Pt film in \nFig. 8 (c) obtained from the temperature evolutions of 〈𝜉〉 and 𝐻𝐾𝑒𝑓𝑓. 〈𝜉〉 varies inversely with \n√𝐻𝐾𝑒𝑓𝑓 in the measured temperature range, which is consistent with the expression 〈𝜉〉 ∝\n1\n𝛼.(𝐻𝐾𝑒𝑓𝑓)1/2. Similarly, we have plotted 〈𝜉〉 as a function of 𝛼𝑇𝑚𝐼𝐺 for the GSGG/TmIG(236 nm)/Pt \nfilm in Fig. 8 (d) obtained from the temperature evolutions of 〈𝜉〉 and 𝛼𝑇𝑚𝐼𝐺 . An inverse \ncorrelation between 〈𝜉〉 and 𝛼𝑇𝑚𝐼𝐺 in the measured temperature range is also apparent from this \nplot, and hence , in agreement with the aforementioned theoretical expression. To establish a more \naccurate correlation between the parameters 〈𝜉〉, 𝐻𝐾𝑒𝑓𝑓and 𝛼, one needs to fix 𝐻𝐾𝑒𝑓𝑓(𝛼), and then \nevaluate 〈𝜉〉 for different values of 𝛼 (𝐻𝐾𝑒𝑓𝑓). It is however challenging to change 𝐻𝐾𝑒𝑓𝑓 of a \nmagnetic material without varying 𝛼 significantly. Nevertheless, we have observed concurrent \nremarkable drops in the LSSE voltage as well as 〈𝜉〉 below 200 K in our TmIG/Pt bilayers \nregardless of TmIG film thickness and substrate choice and correlated the temperature evolution \nof 〈𝜉〉 with the noticeable increases in 𝐻𝐾𝑒𝑓𝑓 and 𝛼 that occur within the same temperature range. \nIt is important to note that FMR probes only the zone center magnons in the GHz range whereas \nthe subthermal magnons which primarily contribute towards the LSSE signal belong to the THz \nregime34. Therefore, the behavior of LSSE cannot be determined by FMR excited by GHz -range \nmicrowaves. As shown by Chang et al.,33 the LSSE voltage varies inversely with 𝛼. Furthermore, \nthe correlation between 〈𝜉〉 and 𝛼 was predicted theoretically18 but never shown experimentally. \nAs indicated in this report, an experimental demonstration of the correlation between 〈𝜉〉, 𝐻𝐾𝑒𝑓𝑓 \nand 𝛼 would be beneficial to fabricate efficient spincaloritronic devices with higher 〈𝜉〉 by tuning \n30 \n these fundamental parameters. However, for deeper understanding of LSSE and 〈𝜉〉, their magnon \nfrequency dependences need to be highlighted . \n \n2. 7. Magnon Frequency Dependence s of the LSSE Voltage and Magnon Propagation Length \n As discussed before, the low energy subthermal magnons with longer 〈𝜉〉 are primarily \nresponsible for LSSE. These low frequency magnons are partially frozen out by the application of \nexternal magnetic field because of increased magnon energy gap due to the Zeeman effect.20,108 \nTherefore, 〈𝜉〉 and hence the LSSE signal is strongly suppressed by the application of high \nmagnetic field.20,108 However, the field induced suppression is dependent on the thickness of the \nmagnetic film.34 If the film thickness is lower than 〈𝜉〉, the low frequency subthermal magnons \ncannot recognize the local temperature gradient and do not participate in LSSE. In that case, only \nhigh frequency magnons with shorter 〈𝜉〉 and much higher energy than the Zeeman energy \ncontribute towards the LSSE signal and hence the field induced suppression of LSSE becomes \nnegligible.20,108 However, if the film thickness is higher than 〈𝜉〉, most of the low frequency \nmagnons contribute towards LSSE and hence, the field induced suppression becomes more \nsignificant.20,108 As we have observed in our TmIG films that the trend of temperature dependent \nLSSE signal is nearly independent of the substrate choice, the temperature dependent 〈𝜉〉 is not \nsupposed to change significantly with the substrate choice . \n31 \n \nFigure 9. Magnetic field induced suppression of the LSSE voltage : 𝑉𝐼𝑆𝐻𝐸(𝐻) loops for the \nsGGG/TmIG(40nm)/Pt nm film at T = (a) 295, (b) 200 and (c) 140 K measured up to high \nmagnetic field of 𝜇0𝐻=9 T. 𝑉𝐼𝑆𝐻𝐸(𝐻) loops for the sGGG/TmIG(75nm)/Pt nm film at T = (d) \n295, (e) 200 and (f) 140 K measured up to high magnetic field of 𝜇0𝐻=9 T. 𝑉𝐿𝑆𝑆𝐸(𝑇,𝜇0𝐻=9T) \nand 𝑉𝐿𝑆𝑆𝐸𝑚𝑎𝑥(𝑇) for (g) 40 nm and (h) 75 nm films. (i) Temperature dependence of 𝛿𝑉𝐿𝑆𝑆𝐸(%) for \n40 nm and 75 nm films. \n \n Therefore, (i) to verify the influence of thickness on the field induced suppression of the \nLSSE signal in TmIG films and (ii) to confirm whether 〈𝜉〉 of the GSGG/TmIG/Pt films obtained \nby analyzing the low field LSSE signal matches closely with that of the sGGG/TmIG/Pt films , we \nperformed the high field LSSE measurements on the sGGG/TmIG/Pt films with thicknesses of 40 \nand 75 nm. Figs. 9 (a)-(c) demonstrate the 𝑉𝐼𝑆𝐻𝐸(𝐻) loops for the 40 nm film at T = 295, 200 and \n140 K measured up to high magnetic field of 𝜇0𝐻=9 T. It can be seen that the LSSE signal for \nthe 40 nm film does not show prominent suppression at 9 T at 295 K . However, as temperature \ndecreases below 200 K, the suppression of the LSSE signal becomes noticeable. On the other hand, \n \n32 \n as seen in Figs. 9 (d)-(f), the LSSE signal for the 75 nm film shows significant suppression even at \n295 K and the suppression of LSSE signal enhances with decreasing temperature. The more intense \nsuppression of the LSSE signal in the 75 nm film compared to the 40 nm film at all temperatures \nbetween 295 and 140 K is also evident from 𝑉𝐿𝑆𝑆𝐸(𝑇) for these two films shown in Figs. 9 (g) and \n(h). We have also estimated the percentage change in 𝑉𝐿𝑆𝑆𝐸 by the application of 9 T magnetic \nfield, which we define as, 𝛿𝑉𝐿𝑆𝑆𝐸(%)=[𝑉𝐿𝑆𝑆𝐸(9 T)−𝑉𝐿𝑆𝑆𝐸𝑚𝑎𝑥\n𝑉𝐿𝑆𝑆𝐸𝑚𝑎𝑥]×100% , where, 𝑉𝐿𝑆𝑆𝐸(9 T) is the \nabsolute value of 𝑉𝐿𝑆𝑆𝐸 at 9 T magnetic field and 𝑉𝐿𝑆𝑆𝐸𝑚𝑎𝑥 is the value of 𝑉𝐿𝑆𝑆𝐸 at the maximum point \nof the 𝑉𝐼𝑆𝐻𝐸(𝐻) loop. As shown in Figs. 9 (i), |𝛿𝑉𝐿𝑆𝑆𝐸| for the 75 nm film is nearly 14% at 295 K \nbut increases to ≈ 32% at 140 K. On the other hand, |𝛿𝑉𝐿𝑆𝑆𝐸| for the 40 nm film is negligible at \n295 K but increases to ≈ 7% at 140 K. These results indicate that 〈𝜉〉 for the sGGG/TmIG/Pt films \nat 295 K is between 40 and 75 nm, which is close to the value of 〈𝜉〉 obtained for the \nGSGG/TmIG/Pt films. Since 〈𝜉〉 decreases at low temperatures and becomes smaller than 40 nm \nbelow 150 K, the sGGG/TmIG(40nm)/Pt film shows significant field induced suppression of 𝑉𝐿𝑆𝑆𝐸 \nat low temperatures. Similarly, since 〈𝜉〉 at low temperatures is much smaller than 75 nm, the field \ninduced suppression of 𝑉𝐿𝑆𝑆𝐸 is also large at low temperatures for the 75 nm film. Note that in case \nof YIG/Pt films, the magnetic field induced suppression of the LSSE signal diminishes with \ndecreasing temperature,34 whereas, an opposite trend has been observed in case of TmIG . Such \nbehavior can be explained by different trends of the temperature dependent 〈𝜉〉 in YIG and TmIG. \nAs explained by Guo et al.,34 the temperature induced enhancement of 〈𝜉〉 neutralizes the field \ninduced suppression of 〈𝜉〉, and because of these two competing factors, the field induced \nsuppression of the LSSE voltage is less prominent at low temperatures in YIG/Pt films. On the \ncontrary, the combined effects of the temperature induced reduction in 〈𝜉〉 observed in our \n33 \n TmIG/Pt films and field induced suppression of 〈𝜉〉 give rise to stronger field induced suppression \nof the LSSE voltage at lower temperatures. \n \nFigure 10. Magnon frequency dispersion for TmIG : (a) Magnon frequency dispersion for TmIG \nfor 𝜇0𝐻=0 T and 9 T magnetic fields at T = 295 K. (b) Comparison of the magnon frequency \ndispersion for YIG and TmIG at room temperature for 𝜇0𝐻=0 T. Magnon frequency dispersion \nof TmIG at different temperatures for (c) 𝜇0𝐻=0 T and (d) 9 T magnetic fields . \n \nNext, to have a qualitative understanding of the magnon frequency dependences of the \nLSSE signal and 〈𝜉〉, we have estimated the magnon frequency dispersion for TmIG in the absence \nand in presence of high magnetic field of 9T. According to the classical Heisenberg ferromagnet \nmodel for spin waves, the parabolic magnon frequency dispersion at low energies can be expressed \nas,17,20,109 ℏ𝜔𝑘= 𝑔𝑒𝑓𝑓𝜇𝐵𝐻+𝐷𝑆𝑊.𝑘2𝑎02+𝐸𝑎𝑛𝑖(𝐾𝑒𝑓𝑓), where the first term represents the \nZeeman energy gap due to the application of external magnetic field, the second term is associated \n \n34 \n with spin wave stiffness (𝐷𝑆𝑊 is the spin wave stiffness constant), and the third term represents \nthe contribution of effective magnetic anisotropy energy, 𝐸𝑎𝑛𝑖(𝐾𝑒𝑓𝑓). Here, the value of 𝐷𝑆𝑊 is \ntaken as that of YIG, i.e., 𝐷𝑆𝑊𝑎02=4.2 ×10−29 erg.cm2 at room temperature.108 Using the \nexpression 𝐾𝑒𝑓𝑓=𝐾𝑠ℎ𝑎𝑝𝑒+𝐾𝑚𝑐 + 𝐾𝑚𝑒, we determined the temperature dependence of \n𝐸𝑎𝑛𝑖(𝐾𝑒𝑓𝑓) for TmIG. Here, the temperature variation of 𝐾𝑚𝑒 was obtained from the temperature \ndependence of 𝜆111 reported in the literature83 (see Supplementary Figures 8(e) ). Since the shear \nmodulus, 𝑐44 in REIGs is weakly dependent on the rare -earth species, the value of 𝑐44 is taken as \nthat of YIG (76.4 GPa at room temperature).47 The temperature dependence of 𝐾𝑠ℎ𝑎𝑝𝑒 was \nestimated from the temperature variation of 𝑀𝑆 (see Supplementary Figures 8(f) ). We assumed \nconstant value of 𝐾𝑚𝑐=0.058 kJ/m3 throughout the measured temperature range.6 As shown in \nSupplementary Figures 8(f), 𝐾𝑒𝑓𝑓 is positive for 𝑇≤300 K (𝐾𝑒𝑓𝑓=20 kJ/m3) and its absolute \nvalue increases considerably with decreasing temperature. Fig. 10 (a) shows the magnon frequency \ndispersion for TmIG for 𝜇0𝐻=0 T and 9 T magnetic fields at T = 295 K. Clearly, the high \nmagnetic field opens a magnon energy gap in the low frequency regime (much smaller than \nthermal energy at room temperature) indicating the suppression of 〈𝜉〉 and hence significant \nreduction of the LSSE signal at high magnetic fields. As shown in Fig. 10 (b), we have compared \nthe magnon frequency dispersion for YIG and TmIG at room temperature for 𝜇0𝐻=0 T. It is \nevident that the opening of magnon energy gap in TmIG is higher than in YIG even in the absence \nof external magnetic field, which is mainly caused by the effective magnetic anisotropy. Note that \n𝐾𝑒𝑓𝑓 of TmIG is higher than that of YIG.110 The higher value of magnon energy gap in TmIG \ncompared to YIG thus indicates the higher possibility of freezing out of the low energy subthermal \nmagnons in TmIG . This, along with higher value of 𝛼 in TmIG contributes to the lower value of \n〈𝜉〉 and hence the LSSE voltage in TmIG compared to YIG16. Furthermore, the value of 𝑔𝑒𝑓𝑓 at \n35 \n room temperature is lower in TmIG (≈1.63)93 than in YIG (≈ 2.046)111. Therefore, for a given \napplied magnetic field strength, the magnitude of the magnon energy gap due to Zeeman effect \nwill be different in TmIG than in YIG. Moreover, 𝛼 in TmIG6,96 is nearly two orders of magnitude \nhigher than in YIG112. In other words, different values of 𝐾𝑒𝑓𝑓, 𝑔𝑒𝑓𝑓 and 𝛼 as well as their different \ntemperature dependences give rise to different temperature profiles of 〈𝜉〉 in TmIG and YIG. In \nFigs. 10 (c) and (d), we show the magnon frequency dispersion of TmIG at different temperatures \nfor 𝜇0𝐻=0 T and 9 T magnetic fields, respectively. Clearly, the magnon energy gap increases \nwith decreasing temperature due to enhanced magnetic anisotropy at low temperatures. \nApplication of 9 T magnetic field increases the magnon energy gap further due to Zeeman ef fect. \nThese results help explain the observed decrease in 〈𝜉〉 and enhanced magnetic field induced \nsuppression of the LSSE signal at low temperatures in TmIG.20,108 \n \nWe believe that our findings will attract the attention of the spintronic community for \nfurther exploration of long-range thermo -spin transport in different REIG based magnetic thin \nfilms and heterostructures for tunable spincaloritronic efficiency by manipulating 𝐻𝐾𝑒𝑓𝑓 and 𝛼. For \nexample, 𝐻𝐾𝑒𝑓𝑓 of the REIG thin films grown on piezoelectric substrates can be modulated by \napplying a gate voltage,113 which can eventually influence 〈𝜉〉 and hence the spincaloritronic \nefficiency. Therefore, our study also provides a step towards the development of efficient \nspincaloritronic devices based on voltage controlled LSSE. \n \n3. CONCLUSION \nIn summary, we have performed a comprehensive investigation of the temperature \ndependent LSSE, RF transverse susceptibility, and broadband FMR measurements on TmIG /Pt \n36 \n bilayers grown on different substrates. The decrease in the LSSE volta ge below 200 K independent \nof TmIG film thickness and substrate choice is attribute d to the increases in 𝐻𝐾𝑒𝑓𝑓 and 𝛼 that occur \nwithin the same temperature range. From the TmIG thickness dependence of the LSSE voltage, \nwe determined the temperature dependence of 〈𝜉〉 and highlighted its correlation with the \ntemperature dependent 𝐻𝐾𝑒𝑓𝑓 and 𝛼 in TmIG/Pt bilayers, which will be beneficial for the \ndevelopment of REIG -based spincaloritronic nanodevices . Furthermore, the enhanced suppression \nof the LSSE voltage by the application of high magnetic field at low temperatures together with \nthe temperature evolution of magnon frequency dispersion in TmIG estimated from the \ntemperature dependent 𝐾𝑒𝑓𝑓 and 𝛼 support our observation of the decrement of 〈𝜉〉 at low \ntemperatures in the TmIG/Pt bilayers. \n \n \n \n \n \n \n \n \n \n \n \n \n \n37 \n 4. METHODS \nThin film growth and structural/morphological characterization : Single -crystalline TmIG thin \nfilms were deposited by pulsed laser deposition (PLD), using two different PLD setups. The thin \nfilms were grown epitaxially on different (111) -oriented substrates, including GGG ( Gd3Ga5O12), \nGSGG ( Gd3Sc2Ga3O12), and sGGG ( (Gd 2.6Ca0.4)(Ga 4.1Mg 0.25Zr0.65)O12). Substrates with (111) \norientation are chosen so that the magnetoelastic anisotropy of the TmIG films favors PMA. Using \nthe first PLD setup, films with varying thickness between 28 nm and 236 nm were grown on GGG \nand GSGG substrates. A KrF excimer laser with a wavelength of 248 nm , a fluence of 3 -4 J/cm², \nand a repetition rate of 2 Hz is used. Before the first deposition, t he TmIG target was preablated \ninside the PLD chamber with more than 104 pulses. All substrates were annealed for 8 h at 1250°C \nin oxygen atmosphere prior to the film deposition to provide a high substrate surface quality . \nGrowth conditions were selected to achieve stoichiometric, single -crystalline thin films with a \nsmooth surface of about 0.2 -0.3 nm in root -mean square roughness (RMS). For all films, the \nsubstrate was heated to 595°C during the film deposition , monitored by a thermocouple inside the \nsubstrate holder. The TmIG thin films were grown at a rate of 0.01 − 0.02 nm/s, in the presence of \nan oxygen background atmosphere of 0.05 mbar. After the deposition, the samples were cooled to \nroom temperature at approximately 5 K/min , maintaining the oxygen atmosphere. A layer of 5 nm \nPt was deposited at room temperature ex-situ on the garnet films by DC magnetron sputtering \nusing a shadow mask. The TmIG films were annealed at 400°C for 1 h inside the sputter chamb er \nprior to the Pt deposition to avoid surface contamination114.To complement these samples, TmIG \nfilms with thicknesses 75 and 40 nm were grown on sGGG substrates using a second PLD setup . \nThe laser wavelength was 248 nm at 10 Hz, the fluence 1.3 J/cm2, and the substrate temperature \n38 \n was ~750 ˚C with an oxygen pressure of 0.2 mbar. Samples were cooled at 20 K/min in 0.2 mbar \noxygen. \n \nThe film surface morphology was investigated by atomic force microscopy (AFM), while \nthe structural properties of the thin films were identified by x -ray diffraction (XRD) using \nmonochromatic Cu Kα radiation. The film thickness was evaluated from the Laue oscillations (for \nthe thinne r films) and by spectroscopic ellipsometry. Further, a cross -sectional high resolution \nscanning transmission electron microscopy (HR -STEM) was conducted, using a JEOL NEOARM \nF200 operated at an electron energy of 200 keV. Electron energy loss spectra (EELS) were \nobtained using a GATAN Continuum S EE LS spectrometer. The cross -sectional sample was \nprepared by mechanical dimpling and ion polishing. Interdiffusion between the TmIG film and \nsubstrate is expected to be limited to a depth of order 1 -3 nm41 and its effects are neglected for the \nfilm thicknesses used in this study. See Supplementary Figure 1 (e) for energy dispersive X -ray \nspectroscopy (EDX) using transmission electron microscopy (TEM) performed on the \nGSGG/TmIG(20 5nm) film. \n \nTemperature dependent MFM measurements : Temperature dependent MFM measurements were \nperformed on a Hitachi 5300E system. All measurements were done under high vacuum (P ≤ 10-6 \nTorr). MFM measurements utilized HQ: NSC18/Co -Cr/Al BS tips, which were magnetized out -\nof-plane with respect to the tip surface via a permanent magnet. Films were first magnetized to \ntheir saturation magnetization by being placed in a 1T static magnetic field, in -plane with the film \nsurface. After that AC demagnetization of the film was implemented before init iating the MFM \nscans. After scans were performed, a parabolic background was subtracted, which arises from the \n39 \n film not being completely flat on the sample stage. Then, line artifacts were subtracted before \nfinally applying a small Gaussian averaging/sharpening filter over the whole image. Phase \nstandard deviation was determined by fitting a Gaussian to the image p hase distribution and \nextracting the standard deviation from the fit parameters. \n \nMagnetometry : The magnetic properties of the samples were measured using a superconducting \nquantum interference device - vibrating sample magnetometer (SQUID -VSM) at temperatures \nbetween 10 K and 350 K. A linear background stemming from the paramagnetic substrate was \nthereby subtracted. Due to a trapped remanent field inside the superconducting coils, the measured \nmagnetic field was corrected using a paramagnetic reference sample. Additionally, a polar \nmagneto -optical Kerr effect (MOKE) setup was used to record out -of-plane hysteresis loops at \nroom temperature. The molecular field coefficient ( MFC ) model was a Python -coded version of \nDionne’s model115 using molecular field coefficients57. \n \nLongitudinal spin Seebeck effect measurements : The longitudinal spin Seebeck effect (LSSE) \nwas measured over a broad temperature window of 120 K ≤ T ≤ 295 K using a custom -built setup \nassembled on a universal PPMS sample puck. During the LSSE measurements, the films were \nsandwiched between two copper blocks, as shown in Fig. 3(a). The s ame sample geometry was \nused for all films and the distance between the contact leads on the Pt surface were fixed at Ly = 3 \nmm for all films. A single layer of thin Kapton tape was thermally affixed to the naked surfaces of \nthe top (cold) and bottom (hot) copper blocks. To ensure a good thermal link between the film \nsurface and the Kapton tape (thermally conducting and electrically insulating) attached to the top \nand bottom blocks, cryogenic Apiezon N -grease was used. Additionally, the Kapton tape \n40 \n electrically insulated the cold (hot) blocks from the top (bottom) surface of the films . The \ntemperatures of both these blocks were controlled individually by two separate temperature \ncontrollers (Scientific Instruments Model no. 9700) to achieve an ultra -stable temperature \ndifference ( ∆𝑇) with [∆𝑇]𝐸𝑟𝑟𝑜𝑟 < ± 2 mK. The top block (cold) was thermally anchored to the base \nof the PPMS puck using two molybdenum screws whereas a 4 -mm-thick Teflon block was \nsandwiched between the puck base and the hot block (bottom) to maintain a temperature difference \nof ~ 10 K between the hot block and the PPMS base. A resistive chip -heater (PT -100 RTD sensor) \nand a calibrated Si -diode thermometer (DT-621-HR silicon diode sensor) were attached to each of \nthese blocks to efficiently control and sense the temperature. The heaters and thermometers \nattached to the copper blocks were connected to the temperature controllers in such a manner that \na temperature gradient develops along the + z-direction that generates a temperature difference, ∆𝑇, \nbetween the top (cold) and bottom (hot) copper blocks. For a given temperature gradient, the in -\nplane voltage generated along the y-direction across the Pt layer due to the ISHE ( 𝑉𝐼𝑆𝐻𝐸) was \nrecorded by a Keithley 2182a nanovoltmeter while sweeping an external in -plane DC magnetic \nfield from positive to negative values along the x-direction. The Ohmic contacts for the voltage \nmeasurements were made by electrically anchoring a pair of ultra-thin gold wires (25 µm diameter) \nto the Pt layer by high quality conducting silver paint (SPI Supplies ). \n \nTransverse susceptibility measurements : The temperature evolution of effective magnetic \nanisotropy in the GSGG/TmIG/Pt film was measured by employing a radio frequency (RF) \ntransverse susceptibility (TS) technique using a home -built self -resonant tunnel diode oscillator \n(TDO) circuit with a resonance frequency of 12 MHz and sensitivity of ±10 Hz. A physical \nproperty measurement system (PPMS) was employed as a platform to scan the external DC \n41 \n magnetic field ( HDC) and temperature. Before the TS measurements, the film was mounted inside \nan inductor coil (L), which is a component of an LC tank circuit. The entire tank circuit was placed \noutside the PPMS except the coil , L, which was positioned at the base of the PPMS sample \nchamber using a multi -purpose PPMS probe insert ed in such a manner that the axial RF magnetic \nfield ( HRF) of amplitude ~ 10 Oe produced inside the coil was always parallel to the film surface, \nbut perpendicular to HDC. For the T mIG with IP easy axis, 𝐻𝐷𝐶⊥film surface , whereas for the \nfilms with OOP easy axis, 𝐻𝐷𝐶∥film surface. When the sample is subject to both HRF and HDC, \nthe dynamic susceptibility of the sample changes which in turn changes the inductance of the coil \nand, hence, the resonance frequency of the LC tank circuit. The relative change in the resonance \nfrequency is proportional to the relative change in the transverse susceptibility of the sample. \nTherefore, TS as a function of HDC was acquired by monitoring the shift in the resonance frequency \nof the TDO -oscillator circuit by employing an Agilent frequency counter . \n \nBroadband f erromagnetic resonance measurements : Broadband ferromagnetic resonance \n(FMR) measurements ( 𝑓 = 6-20 GHz) were performed using a broadband FMR spectrometer \n(NanOscTM Phase -FMR Spectrometer , Quantum Design Inc., USA) integrated to a Dynacool \nPPMS. The TmIG film was firmly affixed on the surface of a commercial 200-μm-wide coplanar \nwaveguide (CPW) (also provided by NanOscTM Phase -FMR Spectrometer, Quantum Design Inc., \nUSA ) using Kapton tape . The TmIG films were placed faced down on the CPW so that the CPW \ncan efficiently transmit the MW signal from the RF source over a broad f-range. The role of the \nKapton tape is to electrically insulate the films from the CPW. An in-plane RF mag netic field, 𝐻𝑅𝐹 \nis generated in close vicinity to the CPW. In presence of an appropriate external in-plane DC \nmagnetic field, 𝐻𝐷𝐶 provided by the superconducting magnet of the PPMS applied along the \n42 \n direction of the MW current flowing through the CPW(𝐻𝐷𝐶⊥𝐻𝑅𝐹) and frequency, 𝐻𝑅𝐹 \nresonantly excites the TmIG film. The spectrometer employs lock -in detection and records the \nfield derivative of the power absorbed ( 𝑑𝑃/𝑑𝐻) by the film when it is excited by a microwave \n(MW) electromagnetic field generated by injecting a MW current to the CPW . \n \nACKNOWLEDGEMENTS \nFinancial support by the US Department of Energy, Office of Basic Energy Sciences, Division of \nMaterials Science and Engineering under Award No. DE -FG02 -07ER46438 at USF and by the \nGerman Research Foundation (DFG) within project No. 318592081AL618/37 -1 at U Augsburg \nare gratefully acknowledged. CR acknowledges support of NSF award DMR 1808190 and \n1954606. \n \nCONFLICT OF INTEREST \nThe authors have no conflicts to disclose. \n \nDATA AVAILABILITY \nThe data that support the findings of this study are available from the corresponding author upon \nreasonable request. \n \n \n43 \n REFERENCES \n(1) Shirsath, S. E.; Cazorla, C.; Lu, T.; Zhang, L.; Tay, Y. Y.; Lou, X.; Liu, Y.; Li, S.; Wang, \nD. Interface -Charge Induced Giant Electrocaloric Effect in Lead Free Ferroelectric Thin -\nFilm Bilayers. Nano Lett. 2019 , 20 (2), 1262 –1271. \n(2) Shirsath, S. E.; Wang, D.; Zhang, J.; Morisako, A.; Li, S.; Liu, X. Single -Crystal -like \nTextured Growth of CoFe2O4 Thin Film on an Amorphous Substrate: A Self -Bilayer \nApproach. ACS Appl. Electron. Mater. 2020 , 2 (11), 3650 –3657. \n(3) Lu, Q.; Li, Y.; Peng, B.; Tang, H.; Zhang, Y.; He, Z.; Wang, L.; Li, C.; Su, W.; Yang, Q.; \nothers. Enhancement of the Spin -Mixing Conductance in Co -Fe-B/W Bilayers by \nInterface Engineering. Phys. Rev. Appl. 2019 , 12 (6), 64035. \n(4) Chumak, A. V; Vasyuchka, V. I.; Serga, A. A.; Hillebrands, B. Magnon Spintronics. Nat. \nPhys. 2015 , 11 (6), 453 –461. \n(5) Cornelissen, L. J.; Liu, J.; Duine, R. A.; Youssef, J. Ben; Van Wees, B. J. Long -Distance \nTransport of Magnon Spin Information in a Magnetic Insulator at Room Temperature. \nNat. Phys. 2015 , 11 (12), 1022 –1026. \n(6) Rosenberg, E. R.; Litzius, K.; Shaw, J. M.; Riley, G. A.; Beach, G. S. D.; Nembach, H. T.; \nRoss, C. A. Magnetic Properties and Growth -Induced Anisotropy in Yttrium Thulium Iron \nGarnet Thin Films. Adv. Electron. Mater. 2021 , 7 (10), 2100452. \n(7) Nakayama, H.; Althammer, M.; Chen, Y. -T.; Uchida, K.; Kajiwara, Y.; Kikuchi, D.; \nOhtani, T.; Geprägs, S.; Opel, M.; Takahashi, S.; others. Spin Hall Magnetoresistance \nInduced by a Nonequilibrium Proximity Effect. Phys. Rev. Lett. 2013 , 110 (20), 206601. \n(8) Shao, Q.; Tang, C.; Yu, G.; Navabi, A.; Wu, H.; He, C.; Li, J.; Upadhyaya, P.; Zhang, P.; \nRazavi, S. A.; others. Role of Dimensional Crossover on Spin -Orbit Torque Efficiency in \n44 \n Magnetic Insulator Thin Films. Nat. Commun. 2018 , 9 (1), 1 –7. \n(9) Evelt, M.; Soumah, L.; Rinkevich, A. B.; Demokritov, S. O.; Anane, A.; Cros, V.; \nYoussef, J. Ben; De Loubens, G.; Klein, O.; Bortolotti, P.; others. Emission of Coherent \nPropagating Magnons by Insulator -Based Spin -Orbit -Torque Oscillators. Phys. Rev. Appl. \n2018 , 10 (4), 41002. \n(10) Heinrich, B.; Burrowes, C.; Montoya, E.; Kardasz, B.; Girt, E.; Song, Y. -Y.; Sun, Y.; Wu, \nM. Spin Pumping at the Magnetic Insulator (YIG)/Normal Metal (Au) Interfaces. Phys. \nRev. Lett. 2011 , 107 (6), 66604. \n(11) Uchida, K.; Adachi, H.; Ota, T.; Nakayama, H.; Maekawa, S.; Saitoh, E. Observation of \nLongitudinal Spin -Seebeck Effect in Magnetic Insulators. Appl. Phys. Lett. 2010 , 97 (17), \n172505. \n(12) Chanda, A.; Holzmann, C.; Schulz, N.; Seyd, J.; Albrecht, M.; Phan, M. -H.; Srikanth, H. \nScaling of the Thermally Induced Sign Inversion of Longitudinal Spin Seebeck Effect in a \nCompensated Ferrimagnet: Role of Magnetic Anisotropy. Adv. Funct. Mater. 2022 , 32 \n(9), 2109170. \n(13) Kalappattil, V.; Das, R.; Phan, M. -H.; Srikanth, H. Roles of Bulk and Surface Magnetic \nAnisotropy on the Longitudinal Spin Seebeck Effect of Pt/YIG. Sci. Rep. 2017 , 7 (1), \n13316. \n(14) Uchida, K.; Takahashi, S.; Harii, K.; Ieda, J.; Koshibae, W.; Ando, K.; Maekawa, S.; \nSaitoh, E. Observation of the Spin Seebeck Effect. Nature 2008 , 455 (7214), 778 –781. \n(15) Bauer, G. E. W.; Saitoh, E.; Van Wees, B. J. Spin Caloritronics. Nat. Mater. 2012 , 11 (5), \n391–399. \n(16) Kehlberger, A.; Ritzmann, U.; Hinzke, D.; Guo, E. -J.; Cramer, J.; Jakob, G.; Onbasli, M. \n45 \n C.; Kim, D. H.; Ross, C. A.; Jungfleisch, M. B. Length Scale of the Spin Seebeck Effect. \nPhys. Rev. Lett. 2015 , 115 (9), 96602. \n(17) Ritzmann, U.; Hinzke, D.; Kehlberger, A.; Guo, E. -J.; Kläui, M.; Nowak, U. Magnetic \nField Control of the Spin Seebeck Effect. Phys. Rev. B 2015 , 92 (17), 174411. \n(18) Ritzmann, U.; Hinzke, D.; Nowak, U. Propagation of Thermally Induced Magnonic Spin \nCurrents. Phys. Rev. B 2014 , 89 (2), 24409. \n(19) Gomez -Perez, J. M.; Vélez, S.; Hueso, L. E.; Casanova, F. Differences in the Magnon \nDiffusion Length for Electrically and Thermally Driven Magnon Currents in Y 3 F e 5 O \n12. Phys. Rev. B 2020 , 101 (18), 184420. \n(20) Jin, H.; Boona, S. R.; Yang, Z.; Myers, R. C.; Heremans, J. P. Effect of the Magnon \nDispersion on the Longitudinal Spin Seebeck Effect in Yttrium Iron Garnets. Phys. Rev. B \n2015 , 92 (5), 54436. \n(21) Jamison, J. S.; Yang, Z.; Giles, B. L.; Brangham, J. T.; Wu, G.; Hammel, P. C.; Yang, F.; \nMyers, R. C. Long Lifetime of Thermally Excited Magnons in Bulk Yttrium Iron Garnet. \nPhys. Rev. B 2019 , 100 (13), 134402. \n(22) Lee, S.; Lee, W.; Kikkawa, T.; Le, C. T.; Kang, M.; Kim, G.; Nguyen, A. D.; Kim, Y. S.; \nPark, N.; Saitoh, E. Enhanced Spin Seebeck Effect in Monolayer Tungsten Diselenide \nDue to Strong Spin Current Injection at Interface. Adv. Funct. Mater. 2020 , 30 (35), \n2003192. \n(23) Kalappattil, V.; Geng, R.; Das, R.; Pham, M.; Luong, H.; Nguyen, T.; Popescu, A.; \nWoods, L. M.; Kläui, M.; Srikanth, H. Giant Spin Seebeck Effect through an Interface \nOrganic Semiconductor. Mater. Horizons 2020 , 7 (5), 1413 –1420. \n(24) Lee, W. -Y.; Kang, M. -S.; Kim, G. -S.; Park, N. -W.; Choi, K. -Y.; Le, C. T.; Rashid, M. U.; \n46 \n Saitoh, E.; Kim, Y. S.; Lee, S. -K. Role of Ferromagnetic Monolayer WSe2 Flakes in the \nPt/Y3Fe5O12 Bilayer Structure in the Longitudinal Spin Seebeck Effect. ACS Appl. \nMater. \\& Interfaces 2021 , 13 (13), 15783 –15790. \n(25) Phan, M. -H.; Trinh, M. T.; Eggers, T.; Kalappattil, V.; Uchida, K.; Woods, L. M.; \nTerrones, M. A Perspective on Two -Dimensional van Der Waals Opto -Spin-Caloritronics. \nAppl. Phys. Lett. 2021 , 119 (25), 250501. https://doi.org/10.1063/5.0069088. \n(26) Lee, W. -Y.; Kang, M. -S.; Kim, G. -S.; Park, N. -W.; Choi, J. W.; Saitoh, E.; Lee, S. -K. \nAsymmetric In -Plane Temperature Contribution in Longitudinal Spin Seebeck Effect \nMeasurements in the Pt/WSe2/YIG Hybrid Structure. J. Phys. Chem. C 2021 , 125 (23), \n13059 –13066. \n(27) Lee, W. -Y.; Park, N. -W.; Kim, G. -S.; Kang, M. -S.; Choi, J. W.; Choi, K. -Y.; Jang, H. W.; \nSaitoh, E.; Lee, S. -K. Enhanced Spin Seebeck Thermopower in Pt/Holey \nMoS2/Y3Fe5O12 Hybrid Structure. Nano Lett. 2020 , 21 (1), 189 –196. \n(28) Lee, W. -Y.; Park, N. -W.; Kang, M. -S.; Kim, G. -S.; Yoon, Y. -G.; Lee, S.; Choi, K. -Y.; \nKim, K. S.; Kim, J. -H.; Seong, M. -J.; others. Extrinsic Surface Magnetic Anisotropy \nContribution in Pt/Y3Fe5O12 Interface in Longitudinal Spin Seebeck Effect by Graphene \nInterlayer. ACS Appl. Mater. \\& Interfaces 2021 , 13 (37), 45097 –45104. \n(29) Kikuchi, D.; Ishida, M.; Uchida, K.; Qiu, Z.; Murakami, T.; Saitoh, E. Enhancement of \nSpin-Seebeck Effect by Inserting Ultra -Thin Fe70Cu30 Interlayer. Appl. Phys. Lett. 2015 , \n106 (8), 82401. \n(30) Yuasa, H.; Tamae, K.; Onizuka, N. Spin Mixing Conductance Enhancement by Increasing \nMagnetic Density. AIP Adv. 2017 , 7 (5), 55928. \n(31) Yun, S. J.; Duong, D. L.; Ha, D. M.; Singh, K.; Phan, T. L.; Choi, W.; Kim, Y. -M.; Lee, \n47 \n Y. H. Ferromagnetic Order at Room Temperature in Monolayer WSe2 Semiconductor via \nVanadium Dopant. Adv. Sci. 2020 , 7 (9), 1903076. \n(32) Li, Y.; Zheng, D.; Fang, B.; Liu, C.; Zhang, C.; Chen, A.; Ma, Y.; Shen, K.; Liu, H.; \nManchon, A.; others. Unconventional Spin Pumping and Magnetic Damping in an \nInsulating Compensated Ferrimagnet. Adv. Mater. 2022 , 34 (24), 2200019. \n(33) Chang, H.; Praveen Janantha, P. A.; Ding, J.; Liu, T.; Cline, K.; Gelfand, J. N.; Li, W.; \nMarconi, M. C.; Wu, M. Role of Damping in Spin Seebeck Effect in Yttrium Iron Garnet \nThin Films. Sci. Adv. 2017 , 3 (4), e1601614. \n(34) Guo, E. -J.; Cramer, J.; Kehlberger, A.; Ferguson, C. A.; MacLaren, D. A.; Jakob, G.; \nKläui, M. Influence of Thickness and Interface on the Low -Temperature Enhancement of \nthe Spin Seebeck Effect in YIG Films. Phys. Rev. X 2016 , 6 (3), 31012. \n(35) Cornelissen, L. J.; Shan, J.; Van Wees, B. J. Temperature Dependence of the Magnon \nSpin Diffusion Length and Magnon Spin Conductivity in the Magnetic Insulator Yttrium \nIron Garnet. Phys. Rev. B 2016 , 94 (18), 180402. \n(36) Geprägs, S.; Kehlberger, A.; Della Coletta, F.; Qiu, Z.; Guo, E. -J.; Schulz, T.; Mix, C.; \nMeyer, S.; Kamra, A.; Althammer, M. Origin of the Spin Seebeck Effect in Compensated \nFerrimagnets. Nat. Commun. 2016 , 7 (1), 10452. \n(37) Yang, B.; Xia, S. Y.; Zhao, H.; Liu, G.; Du, J.; Shen, K.; Qiu, Z.; Wu, D. Revealing \nThermally Driven Distortion of Magnon Dispersion by Spin Seebeck Effect in Gd 3 Fe 5 \nO 12. Phys. Rev. B 2021 , 103 (5), 54411. \n(38) Li, Y.; Zheng, D.; Liu, C.; Zhang, C.; Fang, B.; Chen, A.; Ma, Y.; Manchon, A.; Zhang, \nX. Current -Induced Magnetization Switching across a Nearly Room -Temperature \nCompensation Point in an Insulating Compensated Ferrimagnet. ACS Nano 2022 , 16 (5), \n48 \n 8181 –8189. \n(39) Tang, C.; Sellappan, P.; Liu, Y.; Xu, Y.; Garay, J. E.; Shi, J. Anomalous Hall Hysteresis \nin T m 3 F e 5 O 12/Pt with Strain -Induced Perpendicular Magnetic Anisotropy. Phys. \nRev. B 2016 , 94 (14), 140403. \n(40) Ding, S.; Ross, A.; Lebrun, R.; Becker, S.; Lee, K.; Boventer, I.; Das, S.; Kurokawa, Y.; \nGupta, S.; Yang, J.; others. Interfacial Dzyaloshinskii -Moriya Interaction and Chiral \nMagnetic Textures in a Ferrimagnetic Insulator. Phys. Rev. B 2019 , 100 (10), 100406. \n(41) Caretta, L.; Rosenberg, E.; Büttner, F.; Fakhrul, T.; Gargiani, P.; Valvidares, M.; Chen, \nZ.; Reddy, P.; Muller, D. A.; Ross, C. A.; others. Interfacial Dzyaloshinskii -Moriya \nInteraction Arising from Rare -Earth Orbital Magnetism in Insulating Magnetic Oxides. \nNat. Commun. 2020 , 11 (1), 1 –9. \n(42) Avci, C. O.; Quindeau, A.; Pai, C. -F.; Mann, M.; Caretta, L.; Tang, A. S.; Onbasli, M. C.; \nRoss, C. A.; Beach, G. S. D. Current -Induced Switching in a Magnetic Insulator. Nat. \nMater. 2017 , 16 (3), 309 –314. \n(43) Avci, C. O.; Rosenberg, E.; Caretta, L.; Büttner, F.; Mann, M.; Marcus, C.; Bono, D.; \nRoss, C. A.; Beach, G. S. D. Interface -Driven Chiral Magnetism and Current -Driven \nDomain Walls in Insulating Magnetic Garnets. Nat. Nanotechnol. 2019 , 14 (6), 561 –566. \n(44) Nunley, T. N.; Guo, S.; Chang, L. -J.; Lujan, D.; Choe, J.; Lee, S. -F.; Yang, F.; Li, X. \nQuantifying Spin Hall Topological Hall Effect in Ultrathin Tm 3 Fe 5 O 12/Pt Bilayers. \nPhys. Rev. B 2022 , 106 (1), 14415. \n(45) Shao, Q.; Liu, Y.; Yu, G.; Kim, S. K.; Che, X.; Tang, C.; He, Q. L.; Tserkovnyak, Y.; Shi, \nJ.; Wang, K. L. Topological Hall Effect at above Room Temperature in Heterostructures \nComposed of a Magnetic Insulator and a Heavy Metal. Nat. Electron. 2019 , 2 (5), 182 – \n49 \n 186. \n(46) Vilela, G. L. S.; Abrao, J. E.; Santos, E.; Yao, Y.; Mendes, J. B. S.; Rodr \\’\\iguez -Suárez, \nR. L.; Rezende, S. M.; Han, W.; Azevedo, A.; Moodera, J. S. Magnon -Mediated Spin \nCurrents in Tm3Fe5O12/Pt with Perpendicular Magnetic Anisotropy. Appl. Phys. Lett. \n2020 , 117 (12), 122412. \n(47) Quindeau, A.; Avci, C. O.; Liu, W.; Sun, C.; Mann, M.; Tang, A. S.; Onbasli, M. C.; \nBono, D.; Voyles, P. M.; Xu, Y. Tm3Fe5O12/Pt Heterostructures with Perpendicular \nMagnetic Anisotropy for Spintronic Applications. Adv. Electron. Mater. 2017 , 3 (1), \n1600376. \n(48) Geller, S.; Remeika, J. P.; Sherwood, R. C.; Williams, H. J.; Espinosa, G. P. Magnetic \nStudy of the Heavier Rare -Earth Iron Garnets. Phys. Rev. 1965 , 137 (3A), A1034. \n(49) Holzmann, C.; Ullrich, A.; Ciubotariu, O. -T.; Albrecht, M. Stress -Induced Magnetic \nProperties of Gadolinium Iron Garnet Nanoscale -Thin Films: Implications for Spintronic \nDevices. ACS Appl. Nano Mater. 2022 , 5 (1), 1023 –1033. \n(50) Guo, C. Y.; Wan, C. H.; Zhao, M. K.; Wu, H.; Fang, C.; Yan, Z. R.; Feng, J. F.; Liu, H. \nF.; Han, X. F. Spin -Orbit Torque Switching in Perpendicular Y3Fe5O12/Pt Bilayer. Appl. \nPhys. Lett. 2019 , 114 (19). \n(51) Song, D.; Ma, L.; Zhou, S.; Zhu, J. Oxygen Deficiency Induced Deterioration in \nMicrostructure and Magnetic Properties at Y3Fe5O12/Pt Interface. Appl. Phys. Lett. 2015 , \n107 (4). \n(52) Rezende, S. M.; Rodríguez -Suárez, R. L.; Cunha, R. O.; Rodrigues, A. R.; Machado, F. L. \nA.; Guerra, G. A. F.; Ortiz, J. C. L.; Azevedo, A. Magnon Spin -Current Theory for the \nLongitudinal Spin -Seebeck Effect. Phys. Rev. B 2014 , 89 (1), 14416. \n50 \n (53) Xiao, J.; Bauer, G. E. W.; Uchida, K.; Saitoh, E.; Maekawa, S. Theory of Magnon -Driven \nSpin Seebeck Effect. Phys. Rev. B 2010 , 81 (21), 214418. \n(54) Arana, M.; Gamino, M.; Silva, E. F.; Barthem, V.; Givord, D.; Azevedo, A.; Rezende, S. \nM. Spin to Charge Current Conversion by the Inverse Spin Hall Effect in the Metallic \nAntiferromagnet M n 2 Au at Room Temperature. Phys. Rev. B 2018 , 98 (14), 144431. \n(55) Azevedo, A.; Vilela -Leão, L. H.; Rodríguez -Suárez, R. L.; Santos, A. F. L.; Rezende, S. \nM. Spin Pumping and Anisotropic Magnetoresistance Voltages in Magnetic Bilayers: \nTheory and Experiment. Phys. Rev. B 2011 , 83 (14), 144402. \n(56) Ding, S.; Liang, Z.; Yun, C.; Wu, R.; Xue, M.; Lin, Z.; Ross, A.; Becker, S.; Yang, W.; \nMa, X.; others. Anomalous Hall Effect in Magnetic Insulator Heterostructures: \nContributions from Spin -Hall and Magnetic -Proximity Effects. Phys. Rev. B 2021 , 104 \n(22), 224410. \n(57) Dionne, G. F. Magnetic Oxides ; Springer, 2009; Vol. 14. \n(58) Rosenberg, E.; Bauer, J.; Cho, E.; Kumar, A.; Pelliciari, J.; Occhialini, C. A.; Ning, S.; \nKaczmarek, A.; Rosenberg, R.; Freeland, J. W.; others. Revealing Site Occupancy in a \nComplex Oxide: Terbium Iron Garnet. Small 2023 , 2300824. \n(59) Gross, M. J.; Su, T.; Bauer, J. J.; Ross, C. A. Molecular Field Coefficient Modeling of \nTemperature -Dependent Ferrimagnetism in a Complex Oxide. Press. Phys. Rev. Appl. \n2023 . \n(60) Ciubotariu, O.; Semisalova, A.; Lenz, K.; Albrecht, M. Strain -Induced Perpendicular \nMagnetic Anisotropy and Gilbert Damping of Tm 3 Fe 5 O 12 Thin Films. Sci. Rep. 2019 , \n9 (1), 17474. \n(61) Rosenberg, E. R.; Beran, L.; Avci, C. O.; Zeledon, C.; Song, B.; Gonzalez -Fuentes, C.; \n51 \n Mendil, J.; Gambardella, P.; Veis, M.; Garcia, C.; others. Magnetism and Spin Transport \nin Rare -Earth -Rich Epitaxial Terbium and Europium Iron Garnet Films. Phys. Rev. Mater. \n2018 , 2 (9), 94405. \n(62) Uchida, K.; Ohe, J.; Kikkawa, T.; Daimon, S.; Hou, D.; Qiu, Z.; Saitoh, E. Intrinsic \nSurface Magnetic Anisotropy in Y 3 Fe 5 O 12 as the Origin of Low -Magnetic -Field \nBehavior of the Spin Seebeck Effect. Phys. Rev. B 2015 , 92 (1), 14415. \n(63) Bougiatioti, P.; Klewe, C.; Meier, D.; Manos, O.; Kuschel, O.; Wollschläger, J.; \nBouchenoire, L.; Brown, S. D.; Schmalhorst, J. -M.; Reiss, G. Quantitative \nDisentanglement of the Spin Seebeck, Proximity -Induced, and Ferromagnetic -Induced \nAnomalous Nern st Effect in Normal -Metal –Ferromagnet Bilayers. Phys. Rev. Lett. 2017 , \n119 (22), 227205. \n(64) Kikkawa, T.; Uchida, K.; Shiomi, Y.; Qiu, Z.; Hou, D.; Tian, D.; Nakayama, H.; Jin, X. -\nF.; Saitoh, E. Longitudinal Spin Seebeck Effect Free from the Proximity Nernst Effect. \nPhys. Rev. Lett. 2013 , 110 (6), 67207. \n(65) Ramos, R.; Kikkawa, T.; Uchida, K.; Adachi, H.; Lucas, I.; Aguirre, M. H.; Algarabel, P.; \nMorellón, L.; Maekawa, S.; Saitoh, E. Observation of the Spin Seebeck Effect in Epitaxial \nFe3O4 Thin Films. Appl. Phys. Lett. 2013 , 102 (7), 72413. \n(66) Chanda, A.; DeTellem, D.; Hai Pham, Y. T.; Shoup, J. E.; Duong, A. T.; Das, R.; Cho, S.; \nVoronine, D. V; Trinh, M. T.; Arena, D. A.; others. Spin Seebeck Effect in Iron Oxide \nThin Films: Effects of Phase Transition, Phase Coexistence, and Surface Magn etism. ACS \nAppl. Mater. Interfaces 2022 , 14 (11), 13468 –13479. \n(67) Iguchi, R.; Uchida, K.; Daimon, S.; Saitoh, E. Concomitant Enhancement of the \nLongitudinal Spin Seebeck Effect and the Thermal Conductivity in a Pt/YIG/Pt System at \n52 \n Low Temperatures. Phys. Rev. B 2017 , 95 (17), 174401. \n(68) Jiménez -Cavero, P.; Lucas, I.; Bugallo, D.; López -Bueno, C.; Ramos, R.; Algarabel, P. \nA.; Ibarra, M. R.; Rivadulla, F.; Morellón, L. Quantification of the Interfacial and Bulk \nContributions to the Longitudinal Spin Seebeck Effect. Appl. Phys. Lett. 2021 , 118 (9), \n92404. \n(69) Chanda, A.; Rani, D.; DeTellem, D.; Alzahrani, N.; Arena, D. A.; Witanachchi, S.; \nChatterjee, R.; Phan, M. -H.; Srikanth, H. Large Thermo -Spin Effects in Heusler Alloy -\nBased Spin Gapless Semiconductor Thin Films. ACS Appl. Mater. \\& Interfaces 2023 . \n(70) Henderson Jr, A. J.; Onn, D. G.; Meyer, H.; Remeika, J. P. Calorimetric Study of Yttrium \nand Rare -Earth Iron Garnets between 0.4 and 4.5 K. Phys. Rev. 1969 , 185 (3), 1218. \n(71) Wang, B. S.; Jiang, H. H.; Zhang, Q. L.; Yin, S. T. Thermal Conductivity of Garnet Laser \nCrystals. In High -Power Lasers and Applications IV ; 2008; Vol. 6823, pp 336 –344. \n(72) Prakash, A.; Flebus, B.; Brangham, J.; Yang, F.; Tserkovnyak, Y.; Heremans, J. P. \nEvidence for the Role of the Magnon Energy Relaxation Length in the Spin Seebeck \nEffect. Phys. Rev. B 2018 , 97 (2), 20408. \n(73) Angeles, F.; Sun, Q.; Ortiz, V. H.; Shi, J.; Li, C.; Wilson, R. B. Interfacial Thermal \nTransport in Spin Caloritronic Material Systems. Phys. Rev. Mater. 2021 , 5 (11), 114403. \n(74) Uchida, K.; Kikkawa, T.; Miura, A.; Shiomi, J.; Saitoh, E. Quantitative Temperature \nDependence of Longitudinal Spin Seebeck Effect at High Temperatures. Phys. Rev. X \n2014 , 4 (4), 41023. \n(75) Ashworth, T.; Loomer, J. E.; Kreitman, M. M. Thermal Conductivity of Nylons and \nApiezon Greases. In Advances in Cryogenic Engineering ; Springer, 1973; pp 271 –279. \n(76) Chanda, A.; Rani, D.; Nag, J.; Alam, A.; Suresh, K. G.; Phan, M. H.; Srikanth, H. \n53 \n Emergence of Asymmetric Skew -Scattering Dominated Anomalous Nernst Effect in the \nSpin Gapless Semiconductors Co 1+ x Fe 1 - x CrGa. Phys. Rev. B 2022 , 106 (13), \n134416. \n(77) Venkat, G.; Cox, C. D. W.; Voneshen, D.; Caruana, A. J.; Piovano, A.; Cropper, M. D.; \nMorrison, K. Magnon Diffusion Lengths in Bulk and Thin Film Fe 3 O 4 for Spin Seebeck \nApplications. Phys. Rev. Mater. 2020 , 4 (7), 75402. \n(78) Aharoni, A.; Frei, E. H.; Shtrikman, S.; Treves, D. The Reversible Susceptibility Tensor \nof the Stoner -Wohlfarth Model. Bull. Res. Counc. Isr. 1957 , 6, 215 –238. \n(79) Chanda, A.; Shoup, J. E.; Schulz, N.; Arena, D. A.; Srikanth, H. Tunable Competing \nMagnetic Anisotropies and Spin Reconfigurations in Ferrimagnetic Fe 100 - x Gd x Alloy \nFilms. Phys. Rev. B 2021 , 104 (9), 94404. \n(80) Harder, M.; Cao, Z. X.; Gui, Y. S.; Fan, X. L.; Hu, C. -M. Analysis of the Line Shape of \nElectrically Detected Ferromagnetic Resonance. Phys. Rev. B 2011 , 84 (5), 54423. \n(81) Pearson, R. F. Magnetocrystalline Anisotropy of Rare -Earth Iron Garnets. J. Appl. Phys. \n1962 , 33 (3), 1236 –1242. \n(82) Sayetat, F. Huge Magnetostriction in Tb3Fe5O12, Dy3Fe5O12, Ho3Fe5O12, Er3Fe5O12 \nGarnets. J. Magn. Magn. Mater. 1986 , 58 (3–4), 334 –346. \n(83) Iida, S. Magnetostriction Constants of Rare Earth Iron Garnets. J. Phys. Soc. Japan 1967 , \n22 (5), 1201 –1209. \n(84) Shumate Jr, P. W.; Smith, D. H.; Hagedorn, F. B. The Temperature Dependence of the \nAnisotropy Field and Coercivity in Epitaxial Films of Mixed Rare -Earth Iron Garnets. J. \nAppl. Phys. 1973 , 44 (1), 449 –454. \n(85) Zeng, X. -Y., Lu, X. -J. & Wang, Y. -Q. The Origin of Growth Induced Magnetic \n54 \n Anisotropy in YIG. ACTA Phys. Sin. 1989 , 38, 11. \n(86) Cavé, L.; Al, T.; Loomer, D.; Cogswell, S.; Weaver, L. A STEM/EELS Method for \nMapping Iron Valence Ratios in Oxide Minerals. Micron 2006 , 37 (4), 301 –309. \n(87) Wang, Z. L.; Yin, J. S.; Jiang, Y. D. EELS Analysis of Cation Valence States and Oxygen \nVacancies in Magnetic Oxides. Micron 2000 , 31 (5), 571 –580. \n(88) Tan, H.; Verbeeck, J.; Abakumov, A.; Van Tendeloo, G. Oxidation State and Chemical \nShift Investigation in Transition Metal Oxides by EELS. Ultramicroscopy 2012 , 116, 24–\n33. \n(89) Van Aken, P. A.; Liebscher, B.; Styrsa, V. J. Quantitative Determination of Iron \nOxidation States in Minerals Using Fe L 2, 3 -Edge Electron Energy -Loss near -Edge \nStructure Spectroscopy. Phys. Chem. Miner. 1998 , 25, 323 –327. \n(90) Khurana, B.; Bauer, J. J.; Zhang, P.; Safi, T.; Chou, C. -T.; Hou, J. T.; Fakhrul, T.; Fan, Y.; \nLiu, L.; Ross, C. A. Magnetism and Spin Transport in Platinum/Scandium -Substituted \nTerbium Iron Garnet Heterostructures. Phys. Rev. Mater. 2021 , 5 (8), 84408. \n(91) Dürrenfeld, P.; Gerhard, F.; Chico, J.; Dumas, R. K.; Ranjbar, M.; Bergman, A.; \nBergqvist, L.; Delin, A.; Gould, C.; Molenkamp, L. W.; others. Tunable Damping, \nSaturation Magnetization, and Exchange Stiffness of Half -Heusler NiMnSb Thin Films. \nPhys. Rev. B 2015 , 92 (21), 214424. \n(92) Jermain, C. L.; Aradhya, S. V; Reynolds, N. D.; Buhrman, R. A.; Brangham, J. T.; Page, \nM. R.; Hammel, P. C.; Yang, F. Y.; Ralph, D. C. Increased Low -Temperature Damping in \nYttrium Iron Garnet Thin Films. Phys. Rev. B 2017 , 95 (17), 174411. \n(93) Hellwege, K. H.; Hellwege, A. M. Landolt -Börnstein -Group III, Condensed Matter, Vol. \n12a, Magnetic and Other Properties of Oxides and Related Compounds -Part A: Garnets \n55 \n and Perovskites. Springer, Berlin 1978. \n(94) Crossley, S.; Quindeau, A.; Swartz, A. G.; Rosenberg, E. R.; Beran, L.; Avci, C. O.; \nHikita, Y.; Ross, C. A.; Hwang, H. Y. Ferromagnetic Resonance of Perpendicularly \nMagnetized Tm3Fe5O12/Pt Heterostructures. Appl. Phys. Lett. 2019 , 115 (17), 172402. \n(95) Nembach, H. T.; Silva, T. J.; Shaw, J. M.; Schneider, M. L.; Carey, M. J.; Maat, S.; \nChildress, J. R. Perpendicular Ferromagnetic Resonance Measurements of Damping and \nLand e ́ G- Factor in Sputtered (Co 2 Mn) 1 - x Ge x Thin Films. Phys. Rev. B 2011 , 84 (5), \n54424. \n(96) Wu, C. N.; Tseng, C. C.; Fanchiang, Y. T.; Cheng, C. K.; Lin, K. Y.; Yeh, S. L.; Yang, S. \nR.; Wu, C. T.; Liu, T.; Wu, M.; others. High -Quality Thulium Iron Garnet Films with \nTunable Perpendicular Magnetic Anisotropy by off -Axis Sputtering – Correlatio n between \nMagnetic Properties and Film Strain. Sci. Rep. 2018 , 8 (1), 11087. \n(97) Seiden, P. E. Ferrimagnetic Resonance Relaxation in Rare -Earth Iron Garnets. Phys. Rev. \n1964 , 133 (3A), A728. \n(98) Spencer, E. G.; LeCraw, R. C.; Clogston, A. M. Low -Temperature Line -Width Maximum \nin Yttrium Iron Garnet. Phys. Rev. Lett. 1959 , 3 (1), 32. \n(99) Guo, S.; McCullian, B.; Hammel, P. C.; Yang, F. Low Damping at Few -K Temperatures \nin Y3Fe5O12 Epitaxial Films Isolated from Gd3Ga5O12 Substrate Using a Diamagnetic \nY3Sc2. 5Al2. 5O12 Spacer. J. Magn. Magn. Mater. 2022 , 562, 169795. \n(100) Reichhardt, C.; Reichhardt, C. J. O.; Milošević, M. V. Statics and Dynamics of Skyrmions \nInteracting with Disorder and Nanostructures. Rev. Mod. Phys. 2022 , 94 (3), 35005. \n(101) Ma, X.; Ma, L.; He, P.; Zhao, H. B.; Zhou, S. M.; Lüpke, G. Role of Antisite Disorder on \nIntrinsic Gilbert Damping in L 1 0 FePt Films. Phys. Rev. B 2015 , 91 (1), 14438. \n56 \n (102) Satapathy, S.; Siwach, P. K.; Singh, H. K.; Pant, R. P.; Maurya, K. K. Interfacial Layer \nEffect on the Enhancement of Gilbert Damping in RF Magnetron Sputtered \nY3Fe5O12/Gd3Ga5O12 Thin Films. Phys. B Condens. Matter 2023 , 669, 415278. \n(103) Kambersk \\`y, V. Spin -Orbital Gilbert Damping in Common Magnetic Metals. Phys. Rev. \nB 2007 , 76 (13), 134416. \n(104) Gilbert, T. L. A Phenomenological Theory of Damping in Ferromagnetic Materials. IEEE \nTrans. Magn. 2004 , 40 (6), 3443 –3449. \n(105) Woltersdorf, G.; Kiessling, M.; Meyer, G.; Thiele, J. -U.; Back, C. H. Damping by Slow \nRelaxing Rare Earth Impurities in Ni 80 Fe 20. Phys. Rev. Lett. 2009 , 102 (25), 257602. \n(106) Rebei, A.; Hohlfeld, J. Origin of Increase of Damping in Transition Metals with Rare -\nEarth -Metal Impurities. Phys. Rev. Lett. 2006 , 97 (11), 117601. \n(107) Reidy, S. G.; Cheng, L.; Bailey, W. E. Dopants for Independent Control of Precessional \nFrequency and Damping in Ni 81 Fe 19 (50 Nm) Thin Films. Appl. Phys. Lett. 2003 , 82 \n(8), 1254 –1256. \n(108) Kikkawa, T.; Uchida, K.; Daimon, S.; Qiu, Z.; Shiomi, Y.; Saitoh, E. Critical Suppression \nof Spin Seebeck Effect by Magnetic Fields. Phys. Rev. B 2015 , 92 (6), 64413. \n(109) Shamoto, S.; Yasui, Y.; Matsuura, M.; Akatsu, M.; Kobayashi, Y.; Nemoto, Y.; Ieda, J. \nUltralow -Energy Magnon Anomaly in Yttrium Iron Garnet. Phys. Rev. Res. 2020 , 2 (3), \n33235. \n(110) Zanjani, S. M.; Onbaşlı, M. C. Predicting New Iron Garnet Thin Films with Perpendicular \nMagnetic Anisotropy. J. Magn. Magn. Mater. 2020 , 499, 166108. \n(111) Castel, V.; Vlietstra, N.; Van Wees, B. J.; Youssef, J. Ben. Frequency and Power \nDependence of Spin -Current Emission by Spin Pumping in a Thin -Film YIG/Pt System. \n57 \n Phys. Rev. B 2012 , 86 (13), 134419. \n(112) Sun, Y.; Chang, H.; Kabatek, M.; Song, Y. -Y.; Wang, Z.; Jantz, M.; Schneider, W.; Wu, \nM.; Montoya, E.; Kardasz, B.; others. Damping in Yttrium Iron Garnet Nanoscale Films \nCapped by Platinum. Phys. Rev. Lett. 2013 , 111 (10), 106601. \n(113) Gross, M. J.; Misba, W. A.; Hayashi, K.; Bhattacharya, D.; Gopman, D. B.; Atulasimha, \nJ.; Ross, C. A. Voltage Modulated Magnetic Anisotropy of Rare Earth Iron Garnet Thin \nFilms on a Piezoelectric Substrate. Appl. Phys. Lett. 2022 , 121 (25). \n(114) Jungfleisch, M. B.; Lauer, V.; Neb, R.; Chumak, A. V; Hillebrands, B. Improvement of \nthe Yttrium Iron Garnet/Platinum Interface for Spin Pumping -Based Applications. Appl. \nPhys. Lett. 2013 , 103 (2), 22411. \n(115) Dionne, G. F. Molecular Field Coefficients of Substituted Yttrium Iron Garnets. J. Appl. \nPhys. 1970 , 41 (12), 4874 –4881. \n " }, { "title": "1609.04753v1.Low_damping_sub_10_nm_thin_films_of_lutetium_iron_garnet_grown_by_molecular_beam_epitaxy.pdf", "content": "Low-damping sub-10-nm thin \flms of lutetium iron garnet grown by molecular-beam\nepitaxy\nC. L. Jermain,1,a)H. Paik,1S. V. Aradhya,1R. A. Buhrman,1D. G. Schlom,1, 2and D.\nC. Ralph1, 2\n1)Cornell University, Ithaca, New York 14853, USA\n2)Kavli Institute at Cornell for Nanoscale Science, Ithaca, New York 14853,\nUSA\n(Dated: 9 October 2018)\nWe analyze the structural and magnetic characteristics of (111)-oriented lutetium\niron garnet (Lu 3Fe5O12) \flms grown by molecular-beam epitaxy, for \flms as thin as\n2.8 nm. Thickness-dependent measurements of the in- and out-of-plane ferromagnetic\nresonance allow us to quantify the e\u000bects of two-magnon scattering, along with the\nsurface anisotropy and the saturation magnetization. We achieve e\u000bective damping\ncoe\u000ecients of 11 :1(9)\u000210\u00004for 5.3 nm \flms and 32(3) \u000210\u00004for 2.8 nm \flms,\namong the lowest values reported to date for any insulating ferrimagnetic sample of\ncomparable thickness.\na)Electronic mail: clj72@cornell.edu\n1arXiv:1609.04753v1 [cond-mat.mes-hall] 15 Sep 2016Insulating ferrimagnets are of interest for spintronic applications because they can possess\nvery small damping parameters, as low as 10\u00005in the bulk.1They also provide the potential\nfor improving the e\u000eciency of magnetic manipulation using spin-orbit torques from heavy\nmetals2,3and topological insulators,4,5because ferrimagnetic insulators will not shunt an\napplied charge current away from the material generating the spin-orbit torque. Making\npractical devices from ferrimagnetic insulators will require techniques capable of growing\nvery thin \flms (a few tens of nm and below) while maintaining low damping. Much of the\nprevious research in this \feld has focused on yttrium iron garnet (Y 3Fe5O12, YIG) grown\nby pulsed-laser deposition or o\u000b-axis sputtering,6{10but YIG is just one in a family of rare\nearth iron garnets with potentially useful properties.11Here we examine the magnetic and\nstructural properties of thin, (111)-oriented \flms of lutetium iron garnet (Lu 3Fe5O12, LuIG)\ngrown by an alternative method, molecular-beam epitaxy (MBE).12We \fnd that MBE is\ncapable of providing sub-10-nm \flms with very low values of damping, rivaling or surpassing\nother deposition techniques. We are able to grow LuIG \flms down to 2.8 nm, or 4 layers\nalong the interplanar spacing d111(0.71 nm),11,13while retaining high crystalline quality.\nWe report in- and out-of-plane ferromagnetic resonance measurements as a function of \flm\nthickness, demonstrating reduced two-magnon scattering compared to previous work. We\nachieve e\u000bective damping coe\u000ecients as low as 11 :1(9)\u000210\u00004for 5.3 nm LuIG \flms and\n32(3)\u000210\u00004for 2.8 nm \flms, which can be compared to the best previous report for very\nthin YIG, 38\u000210\u00004for a 4 nm \flm.6\nAs an iron garnet, LuIG has ferrimagnetic properties similar to YIG. The magnetic\nmoments in both materials arise from their Fe3+ions, which interact via super-exchange\nthrough oxygen atoms.11,14In bulk samples, LuIG has a slightly higher room-temperature\nsaturation magnetization (1815 Oe) than YIG (1760 Oe).11,14,15The bulk lattice parameters\nfor LuIG (12.283 \u0017A) and YIG (12.376 \u0017A) di\u000ber by 0.75%.16,17Both materials can be grown\non isostructural gadolinium gallium garnet (Gd 3Ga5O12, GGG) substrates, which have a\ncubic lattice parameter of 12.383 \u0017A. The resulting mismatch causes biaxial tensile strain\nwith a maximum value of 0.81% and 0.07% for LuIG and YIG, respectively. High-quality\nYIG \flms have been grown previously using o\u000b-axis sputter deposition10,18{21and pulsed-\nlaser deposition (PLD).6,22{28The best reported damping values for thin YIG \flms grown\nby PLD to date include 2 :3\u000210\u00004for a 20 nm \flm,63:2\u000210\u00004for a 10 nm \flm treated\nwith a post-growth etching procedure,29and 0:7\u000210\u00004for a 20 nm \flm treated with a post-\n2growth high-temperature anneal.30For o\u000b-axis sputtering, the best reported values include\n6:1\u000210\u00004for a 16 nm \flm,2112:4\u000210\u00004for a 10.2 nm \flm,19and 0:9\u000210\u00004for a 22 nm \flm\nwith a post-growth high-temperature anneal.20Previous measurements of \flms thinner than\n10 nm recorded signi\fcant two-magnon scattering,6,19and much larger damping parameters\nof 38\u000210\u00004for a 4 nm \flm and 16 \u000210\u00004for a 7 nm \flm.6\nHere we report the growth of epitaxial LuIG \flms with thicknesses from 2.8 to 40 nm\nby reactive MBE on (111) GGG substrates. (We study LuIG, rather than YIG, primarily\nbecause Lu is available within our MBE chamber.) Our substrates are prepared by anneal-\ning at 1300 °C for 3 hr in an air furnace to produce well-de\fned unit-cell steps and smooth\nterraces (see Supplementary Information (SI)). During growth, we simultaneously co-supply\nLu and Fe with an accuracy of \u00065%, to achieve the stoichiometric atomic ratio of Lu:Fe=3:5.\nWe use distilled ozone (O 3) at a background pressure of 1 :0\u000210\u00006Torr as the oxidant. The\ngrowth temperature is 950 to 970 °C, achieved by radiatively heating the backside of the\nGGG substrates, which are coated with 400 nm of Pt to enhance thermal absorption. The\nquality of crystal growth is monitored using in-situ re\rection high-energy electron di\u000brac-\ntion (RHEED) along both the [1 \u001610] and [11 \u00162] in-plane azimuthal directions. The RHEED\nintensity oscillations (Fig. 1(a)) indicate layer-by-layer growth,31with an oscillation period\ncorresponding to the d444spacing, which is a quarter of a single LuIG layer ( d111= 0:71 nm)\nalong the (111)-orientation. We also observe sharp RHEED features and clear Kikuchi lines\nduring growth, as seen in Fig. 1(b,c) for a 10 nm \flm, demonstrating that our \flms are of\nhigh crystalline quality. These features are not observed if the \rux drifts more than \u00065%,\nor if the growth temperature is less than 900 °C.\nWe quantify the strain state and verify the crystalline quality with four-circle X-ray\ndi\u000braction (XRD) measurements. The normalized rocking curves for \flms with di\u000berent\nthicknesses (except the 2.8 nm \flm), overlaid in Fig. 2(a), all have full-width at half-\nmaximum (FWHM) values that are less than 0.004 °, limited by the GGG substrate. This\nindicates that our \flms are commensurately strained, and are at the maximal strain state\nof 0.81% set by the lattice mismatch with the substrate. While the rocking curve measure-\nments on the 2.8 nm \flm lack su\u000ecient signal-to-noise for analysis, the thicker \flms suggest\nthat the strain state is also commensurate for this \flm. The surfaces of the \flms are char-\nacterized by atomic force microscopy. Figure 2(b) shows the 2.8 nm \flm, with a measured\nsurface roughness of 0.26 nm (RMS) over a 5 µm x 5 µm scan area. This indicates that the\n3surface quality is substrate limited, which we observe for all thicknesses. Figure 2(c) shows\nthe\u0012=2\u0012XRD patterns of the LuIG thin \flms for all thicknesses grown. The visible Laue\noscillations con\frm thickness measurements we make with the RHEED intensity oscillations\nand \rux calibrations. Low-angle X-ray re\rectively (XRR) determines the \flm thicknesses\nas 2.84(1), 5.33(2), 9.94(2), 20.16(3) and 40.37(10) nm, which we nominally report as 2.8,\n5.3, 10, 20, and 40 nm.\nThe magnetic properties of the MBE-grown LuIG \flms are characterized by measuring\nthe frequency and thickness dependence of ferromagnetic resonance (FMR). The samples\nare placed, LuIG-side down, on a broadband coplanar waveguide so that the Oersted \feld\nof the waveguide excites FMR at GHz frequencies.32We measure the FMR spectra at \fxed\nfrequency by sweeping the applied magnetic \feld, oriented either in-plane (IP) parallel to the\ncoplanar waveguide or out-of-plane (OOP). For the IP measurements, we position the \flm so\nthat the applied magnetic \feld is always along the [11 \u00162] crystal orientation. The measured\nsignal corresponds to the derivative absorption, which we detect via the voltage from a\ndetector diode. We achieve optimal sensitivity using lock-in ampli\fcation by modulating\nboth the input power and the applied \feld. All of the FMR measurements are performed\nat room temperature. Further details of the FMR apparatus are described in the SI.\nFigure 2(d) shows the IP-FMR response at 5 GHz for LuIG samples with di\u000berent thick-\nnesses. Two trends are apparent as the \flm thickness is reduced: (i) the resonance position\nshifts to higher \felds and (ii) the linewidth increases substantially. Below we show that\nboth of these e\u000bects can be explained by two-magnon scattering.33{35We focus \frst on the\nbehavior of the resonance \felds. We have measured the IP-FMR resonances for each \flm\nthickness at frequencies from 1 to 10 GHz. The evolution as a function of frequency is shown\nin Fig. 3(a) and as a function of thickness in Fig. 3(b).\nIn the presence of two-magnon scattering, the IP resonance \feld Hk\nrpredicted by the\nKittel equation in the thin-\flm limit takes the form33,36\nHk\nr(f;t) =s\u00124\u0019M e\u000b(t)\n2\u00132\n\u0000\u00122\u0019f\nj\rj+ \u0001Hr(t)\u00132\n\u00004\u0019M e\u000b(t)\n2;(1)\nwithfthe excitation frequency, tthe \flm thickness, \u0001 Hra renormalization shift associated\nwith two-magnon scattering, and \rthe gyromagnetic ratio. We measured j\rj=2\u0019= 2:77(2)\n4MHz/Oe based on the frequency dependence of the OOP resonance \feld H?\nr(see SI). The\ne\u000bective anisotropy \feld 4 \u0019M e\u000bis expected to depend on the \flm thickness, because it\ncontains contributions from both bulk demagnetization and surface anisotropy:\n4\u0019M e\u000b= 4\u0019Ms+2Ks\nMst: (2)\nHereMsis the saturation magnetization and Ksis the surface anisotropy energy. The renor-\nmalization shift produced by two-magnon scattering can be related to the surface anisotropy\nas33,36\n\u0001Hr(t) =r\u00122Ks\nMst\u00132\n; (3)\nwhereris a parameter characterizing the strength of two-magnon scattering.\nWe performed a global least-squares \ft of Eqs. (1)-(3) to all the data in Fig. 3 using three\n\ftting parameters r, 4\u0019Ms, andKs. As shown by the lines in Fig. 3, we \fnd excellent \fts\nassuming that all three parameters are independent of \flm thickness, obtaining the values\nr= 4:9(2)\u000210\u00004Oe\u00001, 4\u0019Ms= 1609(1) Oe, and Ks=\u00008:52(8)\u000210\u00003erg/cm2. We also\nattempted to \ft the data without the two-magnon contribution (i.e., with the constraint r=\n0 Oe\u00001), but we found signi\fcant discrepancies for the 2.8 \flm, especially at low frequencies\n(see SI). The non-zero value of rimplies that the two-magnon mechanism is active. For our\n2.8 nm \flm, the renormalization shift is \u0001 Hr= 110 Oe, similar to that found in a 2.7 nm\nNiFe \flm.36This is the \frst report of the renormalization shift in iron garnets. The value\nof 4\u0019Msdetermined by the \ft is signi\fcantly lower than the bulk LuIG value of 1815 Oe.15\nThis reduction is qualitatively consistent with the tensile strain in our \flms from the GGG\nsubstrate. The tensile strain is expected to enhance the antiferromagnetic super-exchange\ninteraction between the two inequivalent Fe3+lattices in the LuIG and therefore reduces the\noverall saturation magnetization.11,14The negative sign that we \fnd for Ksindicates that\nthe surface anisotropy reduces the e\u000bective demagnetization \feld 4 \u0019M e\u000bcompared to the\nbulk value. The magnitude of Ksis relatively weak, however (e.g., more than two orders\nof magnitude smaller than Ksfor annealed CoFeB).37With our values for 4 \u0019MsandKs,\nonly for extremely thin LuIG \flms, <0:8 nm, might the magnetic anisotropy be turned\nperpendicular to the sample plane. For any thickness above this, 4 \u0019M e\u000bfavors in-plane\nmagnetization.\n5Next we consider the FWHM linewidths (\u0001 H) of the IP FMR resonances for our LuIG\n\flms as a function of thickness and FMR frequency. The linewidths of our samples are\nsu\u000eciently narrow that small inhomogeneities in the \flms can result in overlapping but dis-\ntinguishable resonances, as has often been seen previously in measurements on thin garnet\n\flms.6,25,38To make an accurate determination of the intrinsic linewidths, we \ft each mea-\nsured curve to the sum of multiple (2 in this analysis) Lorentzian derivative curves with their\nwidths constrained to be identical (see SI for details). This procedure produces values for\nthe linewidth that are consistent with the results for \flms that can be cleaved into samples\nsu\u000eciently small to isolate a single resonance (see SI).\nFigure 4(a) shows the measured frequency dependence of the linewidth for each of our\n\flms. We observe a linear dependence on frequency up to \u00188 GHz. At higher frequencies,\nthe linewidths deviate from linearity, most obviously for the 2.8 and 5.3 nm \flms. This high-\nfrequency curvature is qualitatively consistent with the e\u000bect of two-magnon scattering, as\nobserved previously in PLD-grown YIG \flms.6Using the expression32\n\u0001H(f) =4\u0019\u000bf\nj\rj+ \u0001H0; (4)\nwe can de\fne an e\u000bective Gilbert damping parameter, \u000b, for each value of \flm thickness\nbased on linear \fts to the data below 8 GHz (Fig. 4(b)). The line shown in Fig. 4(b) is a \ft\nto a phenomenological form\n\u000b=\u000bG+\u000b2M\u0012A\nt2+B\nt\u0013\n; (5)\nwith\u000bG= 0:9(6)\u000210\u00004,A\u000b2M= 125(45)\u000210\u00004nm2, andB\u000b2M= 36(11)\u000210\u00004nm.\nOur damping values are among the best reported for any garnet \flm, and for the \frst\ntime extend the viable thickness of low-damping thin \flms well below 10 nm. We measure\n\u000b= 11:1(9)\u000210\u00004for 5.3 nm LuIG \flms and 32(3) \u000210\u00004for 2.8 nm \flms. We speculate\nthat our MBE growth procedure minimizes the amount of surface roughness and other\ndefects even for very thin LuIG \flms, compared to other deposition techniques, and thereby\nprovides a reduced level of two-magnon scattering. Similar MBE growth procedures may\nalso allow the production of sub-10-nm \flms made from YIG and other garnets, assisting in\nthe development of a wide variety of spintronic devices incorporating these materials.\n6ACKNOWLEDGMENTS\nWe acknowledge F. Guo for helpful discussion on two-magnon theory. This work was\nsupported by the National Science Foundation (DMR-1406333) with partial support from\nthe Cornell Center for Materials Research (CCMR), part of the NSF MRSEC program\n(DMR-1120296). We made use of the Cornell Nanoscale Facility, a member of the National\nNanotechnology Coordinated Infrastructure (NNCI), which is supported by the NSF (ECCS-\n1542081) and also the CCMR Shared Facilities. The work of H.P. and D.G.S. is supported\nby the Air Force O\u000ece of Scienti\fc Research under award number FA9550-16-1-0192.\nREFERENCES\n1M. Sparks, Ferromagnetic-Relaxation Theory (McGraw-Hill, New York, 1964).\n2I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V. Costache, S. Au\u000bret,\nS. Bandiera, B. Rodmacq, A. Schuhl, and P. Gambardella, Nature 476, 189 (2011).\n3L. Liu, C.-F. C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, Science\n336, 555 (2012).\n4A. R. Mellnik, J. S. Lee, A. Richardella, J. L. Grab, P. J. Mintun, M. H. Fischer, A. Vaezi,\nA. Manchon, E. A. Kim, N. Samarth, and D. C. Ralph, Nature 511, 449 (2014).\n5Y. Fan, P. Upadhyaya, X. Kou, M. Lang, S. Takei, Z. Wang, J. Tang, L. He, L.-T. Chang,\nM. Montazeri, G. Yu, W. Jiang, T. Nie, R. N. Schwartz, Y. Tserkovnyak, and K. L. Wang,\nNat. Mater. 13, 699 (2014).\n6O. d'Allivy Kelly, A. Anane, R. Bernard, J. Ben Youssef, C. Hahn, A. H. Molpeceres,\nC. Carretero, E. Jacquet, C. Deranlot, P. Bortolotti, R. Lebourgeois, J.-C. Mage,\nG. de Loubens, O. Klein, V. Cros, and A. Fert, Appl. Phys. Lett. 103, 082408 (2013).\n7C. Hahn, V. V. Naletov, G. de Loubens, O. Klein, O. d'Allivy Kelly, A. Anane, R. Bernard,\nE. Jacquet, P. Bortolotti, V. Cros, J. L. Prieto, and M. Mu~ noz, Appl. Phys. Lett. 104,\n152410 (2014).\n8M. Montazeri, P. Upadhyaya, M. C. Onbasli, G. Yu, K. L. Wong, M. Lang, Y. Fan, X. Li,\nP. Khalili Amiri, R. N. Schwartz, C. A. Ross, and K. L. Wang, Nat. Commun. 6, 8958\n(2015).\n9A. Hamadeh, O. d'Allivy Kelly, C. Hahn, H. Meley, R. Bernard, A. H. Molpeceres,\n7V. V. Naletov, M. Viret, A. Anane, V. Cros, S. O. Demokritov, J. L. Prieto, M. Mu~ noz,\nG. de Loubens, and O. Klein, Phys. Rev. Lett. 113, 197203 (2014).\n10H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel, and F. Y. Yang, Phys. Rev. Lett.\n112, 197201 (2014).\n11S. Geller, A. Paoletti, P. Hansen, J. C. Slonczewski, A. P. Malozemo\u000b, P. E. Wigen,\nR. W. Teale, A. Tucciarone, U. Enz, J. F. Dillon, R. Metselaar, P. K. Larsen, G. B. Scott,\nC. Rudowicz, W. Jantz, W. Wettling, J. Schneider, R. Krishnan, and W. Tolksdorf,\nProc. Int'l School Phys.Enrico Fermi, Course LXX , edited by A. Paoletti, Vol. 14 (North-\nHolland Company, Amsterdam, 1978).\n12D. G. Schlom, L. Chen, X. Pan, A. Schmehl, and M. A. Zurbuchen, J. Am. Ceram. Soc.\n91, 2429 (2008).\n13A. Kelly and K. M. Knowles, Crystallography and Crystal Defects (John Wiley & Sons,\nLtd, West Sussex, 2012).\n14E. Anderson, Phys. Rev. 134, A1581 (1964).\n15L. G. Van Uitert, E. M. Gyorgy, W. A. Bonner, W. H. Grodkiewicz, E. J. Heilner, and\nG. J. Zydzik, Mater. Res. Bull. 6, 1185 (1971).\n16S. Geller and M. M. Gilleo, J. Phys. Chem. 3, 30 (1957).\n17G. P. Espinosa, J. Chem. Phys. 37, 2344 (1962).\n18H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel, and F. Y. Yang, Phys. Rev. B\n88, 100406 (2013).\n19T. Liu, H. Chang, V. Vlaminck, Y. Sun, M. Kabatek, A. Ho\u000bmann, L. Deng, and M. Wu,\nJ. Appl. Phys. 115, 17A501 (2014).\n20H. Chang, P. Li, W. Zhang, T. Liu, A. Ho\u000bmann, L. Deng, and M. Wu, IEEE Magn.\nLett.5, 1 (2014).\n21J. T. Brangham, K.-Y. Meng, A. S. Yang, J. C. Gallagher, B. D. Esser, S. P. White, S. Yu,\nD. W. McComb, P. C. Hammel, and F. Yang, Phys. Rev. B 94, 054418 (2016).\n22P. C. Dorsey, S. E. Bushnell, R. G. Seed, and C. Vittoria, J. Appl. Phys. 74, 1242 (1993).\n23S. A. Manuilov, S. I. Khartsev, and A. M. Grishin, J. Appl. Phys. 106, 123917 (2009).\n24S. A. Manuilov and A. M. Grishin, J. Appl. Phys. 108, 013902 (2010).\n25B. Heinrich, C. Burrowes, E. Montoya, B. Kardasz, E. Girt, Y.-Y. Song, Y. Sun, and\nM. Wu, Phys. Rev. Lett. 107, 066604 (2011).\n26M. C. Onbasli, A. Kehlberger, D. H. Kim, G. Jakob, M. Kl aui, A. V. Chumak, B. Hille-\n8brands, and C. A. Ross, APL Mater. 2, 106102 (2014).\n27B. M. Howe, S. Emori, Hyung-Min Jeon, T. M. Oxholm, J. G. Jones, K. Mahalingam, Yan\nZhuang, N. X. Sun, and G. J. Brown, IEEE Magn. Lett. 6, 3500504 (2015).\n28C. Tang, M. Aldosary, Z. Jiang, B. Madon, K. Chan, J. E. Garay, and J. Shi, Appl. Phys.\nLett.108, 102403 (2016).\n29Y. Sun, Y.-Y. Song, H. Chang, M. Kabatek, M. Jantz, W. Schneider, M. Wu,\nH. Schultheiss, and A. Ho\u000bmann, Appl. Phys. Lett. 101, 152405 (2012).\n30C. Hauser, T. Richter, N. Homonnay, C. Eisenschmidt, M. Qaid, H. Deniz, D. Hesse,\nM. Sawicki, S. G. Ebbinghaus, and G. Schmidt, Sci. Rep. 6, 20827 (2016).\n31A. Ichimiya and P. I. Cohen, Re\rection High-Energy Electron Di\u000braction (Cambridge\nUniversity Press, Cambridge, 2004).\n32S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L. Schneider, P. Kabos, T. J. Silva,\nand J. P. Nibarger, J. Appl. Phys. 99, 093909 (2006).\n33R. Arias and D. L. Mills, Phys. Rev. B 60, 7395 (1999).\n34D. L. Mills and S. M. Rezende, Top. Appl. Phys. 87, 27 (2003).\n35K. Lenz, H. Wende, W. Kuch, K. Baberschke, K. Nagy, and A. J\u0013 anossy, Phys. Rev. B\n73, 144424 (2006).\n36A. Azevedo, A. B. Oliveira, F. M. de Aguiar, and S. M. Rezende, Phys. Rev. B 62, 5331\n(2000).\n37D. C. Worledge, G. Hu, D. W. Abraham, J. Z. Sun, P. L. Trouilloud, J. Nowak, S. Brown,\nM. C. Gaidis, E. J. O'Sullivan, and R. P. Robertazzi, Appl. Phys. Lett. 98, 022501 (2011).\n38A. Jalali, S. Kahl, V. Denysenkov, and A. Grishin, Phys. Rev. B 66, 104419 (2002).\n9FIGURES\n10b\na\nc [112] [110]FIG. 1. (a) RHEED intensity oscillations of a 10 nm thick LuIG \flm grown on a (111) GGG\nsubstrate, indicating layer-by-layer growth. Each oscillation peak-to-peak corresponds a single\nd444(d111=4) spacing. (b,c) Kikuchi lines in the RHEED image taken along both [1 \u001610] and [11 \u00162]\nazimuthal directions.\n11(nm)\n0.81.0\n0.6\n0.4\n0.2\n0.0\nc da b\n2 μmFIG. 2. (a) X-ray di\u000braction (XRD) rocking curves for all of the LuIG thicknesses above 2.8 nm\nindicate commensurate growth and consistent strain. (b) Representative atomic force microscopy\nimage of the surface of the 2.8 nm \flm, showing a RMS roughness of 0.26 nm over 5 µm x 5 µm\nscan size, which indicates the roughness is substrate limited. (c) \u0012=2\u0012XRD scans of LuIG thin\n\flms grown on (111) GGG substrates as a function of \flm thickness. The asterisk marks the 444\nGGG substrate re\rection. (d) Normalized derivative-absorption FMR spectra of the correspond-\ning samples taken at 5 GHz show narrow linewidths that decrease for increasing thickness. The\nresonance position also depends on the thickness.\n12a bFIG. 3. (a,b) In-plane FMR resonance \felds of each LuIG sample (a) as a function of frequency for\ndi\u000berent sample thicknesses and (b) as a function of thickness for di\u000berent frequencies. The solid\nlines in (a) and (b) represent simultaneous \fts to Eq. (1) with the 3 \ftting parameters r, 4\u0019Ms,\nandKs.\n1311.1(9)31.5(3.0)\n6.2(4)\n3.4(8)1.8(4)this worka\nbFIG. 4. (a) Frequency dependence of the FMR linewidth, for LuIG \flms of di\u000berent thickness.\nThe linewidths are \ft to straight lines up to 8 GHz, after which the linewidths start to roll o\u000b,\nfollowing the signature of two-magnon scattering. (b) Thickness dependence our measured values\nof magnetic damping (black squares). The line depicts the phenomenological form of Eq. (5).\nPreviously-reported results for damping in thin YIG \flms are shown for \flms deposited by PLD\n(open blue symbols) PLD and o\u000b-axis sputtering (open red symbols). Open triangles represent\npost-processed \flms.\n14" }, { "title": "1703.07515v1.Fast_domain_wall_motion_induced_by_antiferromagnetic_spin_dynamics_at_the_angular_momentum_compensation_temperature_of_ferrimagnets.pdf", "content": "Fast domain wall motion induced by antiferromagnetic spin dynamics at the angular momentum \ncompensation temperature of ferrimagnets \n \nKab-Jin Kim1,2†★, Se Kwon Kim3†, Takayuki Tono1, Se-Hyeok Oh4,Takaya Okuno1, Woo Seung Ham1, \nYuushou Hirata1, Sanghoon Kim1, Gyoungchoon Go5, Yaroslav Tserkovnyak3, Arata Tsukamoto6, \nTakahiro Moriyama1, Kyung -Jin Lee4,5,7★, and Teruo Ono1★ \n \n \n \n1Institute for Chemical Research, Kyoto University, Gokasho, Uji, Kyoto, 611 -0011, Japan \n2 Department of Physics, Korea Advanced Institu te of Science and Technology, Daejeon 34141, Korea \n3Department of Physics and Astronomy, University of California, Los Angeles, California 90095, USA \n4Department of Nano -Semiconductor and Engineering, Korea University, Seoul 02841, Korea \n5Department of Mat erials Science & Engineering, Korea University, Seoul 02841, South Korea \n6College of Science and Technology, Nihon University, Funabashi, Chiba 274 -8501, Japan \n7KU-KIST Graduate School of Converging Science and Technology, Korea University, Seoul 02841, \nSouth Korea \n \n★ Correspondence to: kabjin@scl.kyoto -u.ac.jp, kj_lee@korea.ac.kr , ono@scl.kyoto -u.ac.jp \n Antiferromagnetic spintronics is an emerging research field which aims to utilize \nantiferromagnets as core elements in spintronic devices1,2. A central mo tivation toward this \ndirection is that antiferromagnetic spin dynamics is expected to be much faster than ferromagnetic \ncounterpart because antiferromagnets have higher resonance frequencies than ferromagnets3. \nRecent theories indeed predicted faster dynam ics of antiferromagnetic domain walls (DWs) than \nferromagnetic DWs4-6. However, experimental investigations of antiferromagnetic spin dynamics \nhave remained unexplored mainly because of the immunity of antiferromagnets to magnetic fields. \nFurthermore, this immunity makes field -driven antiferromagnetic DW motion impossible despite \nrich physics of field -driven DW dynamics as proven in ferromagnetic DW studies. Here we show \nthat fast field -driven antiferromagnetic spin dynamics is realized in ferrimagnets at t he angular \nmomentum compensation point TA. Using rare -earth–3d-transition metal ferrimagnetic compounds \nwhere net magnetic moment is nonzero at TA, the field -driven DW mobility remarkably enhance s \nup to 20 km s−1T−1. The collective coordinate approach generalized for ferrimagnets7 and atomistic \nspin model simulations6,8 show that this remarkable enhancement is a consequence of \nantiferromagnetic spin dynamics at TA. Our finding allows us to inve stigate the physic s of \nantiferromagnetic spin dynamics and highlights the importance of tuning of the angular \nmomentum compensation point of ferrimagnets, which could be a key towards ferrimagnetic \nspintronics. Encoding information using magnetic DW motion is essential for future magnetic memory \ndevices , such as racetrack memor ies9,10 . High-speed DW motion is a key prerequisite for making the \nracetrack feasible. However, velocity breakdown due to the angular precession of DW , referred to as the \nWalker breakdown11, generall y limits the functional performance in ferromagnet -based DW devices. \nRecently, it was reported that the DW speed boosts up significantly in antiferromagnets due to the \nsuppression of the angular precession4-6. However, the immunity of antiferromag nets to m agnetic fields \nyields notorious difficulties in creating, manipulating , and detecting antiferromagnetic DW s, com pared to \nferromagnetic one s. One possibility to avoid these difficulties is offered by the synthetic \nantiferromagnets12, where the net magnetic moment can be controlled by tuning the thickness of two \nferromagnetic layers coupled antiferromagnetically. However, they still suffer from the field -immunity \nwhen the net magnetic moment approaches zero , preventing the study of antiferromagnetic DW \ndynami cs. Here we show that magnetic field -controlled antiferromagnetic spin dynamics can be achieved \nby employing ferrimagnets. \nThere are a class of ferrimagnet s, rare earth ( RE)–transition metal ( TM) compound s, where the \nspins of two inequivalent sublattices are coupled antiferromagnetically. Because of different Land é g-\nfactors between RE and TM elements , these ferrimagnet s have two special temperatures below the Curie \ntemperature : the magneti sation compensation temperature, TM, at which the two magnetic momen ts \ncancel each other , and the angular momentum compensa tion temperature , TA, at which the net angular \nmomentum vanishes13-15. In particular, the existence of TA in ferrimagnets provides a framework to \ninvestigate antiferromagnetic spin dynamics. It is beca use the time evolution of the state of a magnet is \ngoverned by the commutation relation of the angular momentum, not of the magnetic moment . As a \nresult, the nature of the dynamics of the ferrimagnets is expected to change from ferromagnetic to \nantiferroma gnetic as approaching the angular momentum compensation point TA. Furthermore, the net \nmagnetic moment of ferrimagnets is nonzero at TA and can thus couple to an external magnetic field , \nopen ing a new possibility of field -driven antiferromagnetic spin dyna mics. In order to pursuit this possibility, w e first describe distinguishing features of ferrimagnetic DWs \nnear TA based on the collective coordinate approach7. A ferrimagnetic DW effectively acts as an \nantiferromagnetic DW around TA. The low -energy dynami cs of a DW in quasi one -dimensional magnets \nis generally described by two collective coordinates, its position X and angle Φ, which capture the \ntranslational and spin -rotational degrees of freedom of the DW , respectively. In ferromagnets, they are \ngyrotropically coupled by the Berry phase that is proportional to the net spin density , and the motion of a \nDW slows down severely above a certain critical stimulus, engendering the phenomenon of the Walker \nbreakdown11. In antiferromagnets, on the other hand, the dynamics of X and Φ are independent owing to \nvanishing the net spin density7. The DW dynamics in antiferromagnets is thus free from the Walker \nbreakdown and can be fast in a broad range of external driving forces compar ed to that in ferromagnets. \nIn the following, we explain a theory for the field-driven DW dynamics in ferrimagnets in high \nfields , which agrees with our experi ment al results as discussed later. The DW velocity can be derived as \nfollows by invoking the energy conservation and the gyrotropic coupling between the two collective \ncoordinates X and Φ (see Supplementary Information for microscopic derivations) . For a sufficiently \nstrong external field , the anisotropy energy can be neglected and the time derivatives of X and Φ can be \nconsidered to approach constant values16, \nV X and \n . The energy -dissipation rate caused by \nthese dynamics is given by \n2 2 2\n22 112 V sαsα P Α , where \nΑ is the cross sectional area of \nthe magnet, \n is the domain -wall width, \n1α and \n1sare the Gilbert damping constant and the spin angular \nmomentum density of one sublattice, respectively, and \n2α and \n2s are for the other sublattice. Invoking the \nconservation of the total energy, the rate of energy dissipation can be equated to the decreasing rate of the \nZeeman energy induced by the translation motion of the domain wall, which yields the equation, \nVHM M V sαsα Α Α2 12 2 2\n22 11 2 2 \n, where \n1M and \n2M are th e magnetization of the \ntwo sublattices, and \nH is an external field. Note that the net magnetization, \n2 1M M , does not vanish at \nTA due to the difference in the Land é g-factors of two sublattice atoms , which is essential to drive a DW with an external field. In addition, when there is finite net spin angular momentum \n2 1ss , e.g., away \nfrom TA, the angular and linear velocities are related by the gyrotropic coupling whose strength is \nproportional to the net angular momentum \n2 1ss17. Balancing the gyrotropic force on Φ with the \ndissipative force yields \nVss sαsα2 1 22 11 . Solving the two aforementioned equations for \nV\nand \n , we obtain \n \n H\nss s s)M M(s sV2\n2 12\n22 112 1 22 11\n \n\n, \n\n H\nss s s)M M(ss\n2\n2 12\n22 112 1 2 1\n\n (1) \nAs the system approaches the angu lar momentum compensation point \n02 1ss , the domain wall \nspeed \nV increases, whereas the precessional frequency \n decreases. At T = TA, X and Φ are completely \ndecoupled and the pure translational dynamics of the DW is obtained , implying that the ferrimagnetic DW \neffectively acts as antiferromagnetic DW and its motion is driven by a magnetic field at T = TA. \nIn order to prove the above theoretical prediction, we investigate DW dynamics in ferrimagneti c \nGdFeCo compounds. Figure 1 shows a schematic illustration of our sample. SiN(5 nm)/Gd 23Fe67.4Co9.6 \n(30 nm)/SiN(5 nm) films are deposited on intrinsic Si substrate by magnetron sputtering . GdFeCo is a \nwell-known RE– TM ferrimagnet ic compound , in which RE and TM moment s are coupled \nantiferromagnetically18. The relative mag netic moment s of RE and TM can be easily controlled by \nvarying the composition or temperature, so that TM and TA can be easily designed in RE –TM \nferrimagnet s. The GdFeCo film is then patte rned into micro wires with 5 μm wi dth and 65 μm length \nusing electron beam lithography and Ar ion milling . A Hall bar is designed to detect the DW motion via \nthe anomalous Hall effect (AHE) voltage , VH. \nWe first characterise the magnetic properties of the GdFeCo microstrips . Figure 2 a shows the \nhysteresis loop s of GdFeCo microstrips at various temperatures . The AHE resistance , RH (RH = VH /I), is \nmeasured by sweeping the out-of-plane magnetic field , BZ. Square hysteresis loops are clearly observed, \nindicating that GdFeCo has a perpend icular magnetic anisotropy. The coercivity field, BC, and the \nmagnitude of the Hall resistance change , ΔRH \n Z Z B R B R R H H H , are extracted from the hysteresis loop s and summarised in Fig. 2b. BC increases with increasing temperature , but a sudden drop \nis observed at T = 220 K. A sign change of RH is observed at the same temperature. This is a typical \nbehaviour of ferrimagnet s at the magnetisation compensation temperature TM19. As T approaches TM, the \nnet magnetic moment converges to zero , and thus , a large r magnetic field is required to obtain a \nsufficiently high Zeeman energy to switch the magnetisati on. Thus, BC diverges at TM. The s ign change of \nRH represents additional evidence of TM. The magneto -transport properties of GdFeCo are known to be \ndominated by FeCo mome nts because the 4 f shell, which is respons ible for the magnetic properties of Gd, \nis located far below the Fermi energy level20. Thus, the sign change of RH indicates a change in the \nrelative direction of the FeCo moments with respect t o the magnetic field, which occur s at TM. At T < TM, \nthe Gd moment dominates over the FeCo moment so that the Gd moment aligns alo ng the magnetic field \ndirection . However, at T > TM, the FeCo moment is dominant and thus aligns along the magnetic field \ndirection. Therefore, Fig. 2b allows us to identify TM for our GdFeCo sample , which is approximately 220 \nK. \nAlthough TM can be easily determined by magnetisation or magneto -transport measurement s, it is \ngenerally not easy to determine TA because TA is not related to the net magnetisation but rather to the \nangular momentum of the system. For GdFeCo, the net magneti sation \nM and angular momentum \nA are \nwritten as follows13,14,21. \nFeCo Gd M M M\n and \nFeCo FeCo Gd Gd FeCo Gd / M / M A AA \n , where \n FeCoGdM\n and \n FeCoGdA are the magnetic moment and angular momentum of the Gd (FeCo) sub -lattices , \nrespectively , and \nB g FeCoGd FeCoGd is the gyromagnetic ratio of Gd (FeCo ), where \nB is the Bohr \nmagneton and \n is the reduced Plank constant. According to the literature s22–24, \nFeCog (~2.2) is slightly \nlarger than \nGdg (~2) owing to the spin -orbit coupling of FeCo and zero orbital angular momentum of the \nhalf-filled 4 f shell of Gd ; therefore, TA is expected to be higher than TM in GdFeCo. \nBased on above consideration , we measure the field-driven DW speed at T > TM using a real-time \nDW detection method25–27. We first saturate the magnetisation by applying a large negative field (| B| > \n|BC|) and then switch the field to the positive direction . Thi s positive field is a DW driving field , Bd, and should be smaller than BC (|Bd| < |BC|). Next , we inject a current pulse into the electrode to create a DW by \na current -induced Oersted field , as shown in Fig. 1. The created DW propagate s along the wire due to the \npresence of Bd, and the DW motion is detected at the Hall bar by monitoring the change s in VH. Here , the \nchange s in VH are recorded by an oscilloscope such that nanosecond time -resolution can be achieved . The \nDW speed can be calculated from the arri val time and the travel distance (60 μm) of the DW. The details \nof the measureme nt scheme are explained in the Method section . \nFigure 3 a shows the DW speed as a function of Bd at several temperatures above TM. The DW \nvelocity increases linearly with field for all temperatures. Such a linear behaviour can be described by\n0 dB B v\n. Here , μ is referred to as the DW mobility and B0 is the correction field, which generally \narises from imperfection s of the sample or complexit ies of the internal DW structure28,29. Figure 3b shows \nthe DW velocity as a function of temperature for several bias fields. The DW velocity shows a sharp peak \nas expected near T = TA based on Eq. (1). The DW mobility μ is estimat ed from the linear fit in Fig. 3 a \n(dashed lines) and is plotted as a fu nction of the temperature in Fig. 3 c. Starting from T = 260 K, which is \nslightly higher than TM, μ increases steeply , reach ing its maximum at T = 310 K, and then decreases with \na further increase in temperature. The peak mobility is as high as 20 km ·s−1·T−1 at T = 310 K. These \nexperimental results are in agreement with the analytical expression Eq. (1), which predicts a Lorentz ian \nshape of the DW velocity near T = TA with the width \n22 11 sαsα~ . Such a consistency between \nexperiment and theory manifests that the ferrimagnetic DW indeed acts as an antiferromagnetic DW at T \n= TA. \nWe next perform atomic spin model simulations based on the atomistic Landau -Lifshitz -Gilbert \n(LLG) equation5,8 (see Method section for details) to verify the experimental result and theor etical \nprediction . We employ a set of the reduced magnetic moments around s1−𝑠2=0 as shown in Table 1. \nThe total number of set is 9 in which the index 5 corresponds to the temperature at TA. We assume that a \nDW is of Bloch -type with perpendicular mag netic anisotropy along the z axis. Fig ure 4a shows the DW \nvelocity as a function of the external field BZ, applied along the z axis. Velocities increase linearly with BZ for BZ > 10 mT . The numerical results (circular symbols) are in excellent agreement w ith the analytic \nsolution (solid lines) for the DW velocity for high fields in Eq. (1). The inset of Fig. 4a shows the DW \nvelocity in low field regime s (BZ < 10 mT) . The Walker breakdown occurs in this regime except for the \ncase of TA. The vertical dashed lines represent the Walker breakdown field. Fig ure 4b shows the DW \nvelocity as a function of δs=s1−𝑠2. At δs=0, the DW velocity is the highest , which agrees with the \nexperimental result and theory . This good agreement also supports that field -driven antiferromagnetic \nspin dynamics is realized in ferrimagnets at T A. \n To date , the angular momentum compensation point and its effect on the magnetisation dynamics \nhave often been overlooked in studies of ferrimagnet s. Laser-induced magnetisation switching18,30,31 and \nmagnetic DW motion32–33, which have bee n major research themes of ferrima gnets, have mostly been \nstudied without identifying TA. It has been investigated in the context of magnetic resonance or \nmagnetisation switching by current around TA 14,15,34, but a clear identification of spin dynamics at T = TA \nhave been remained elusive15. On the other hand, our results clearly show that the antiferromagnetic spin \ndynamics is achievable at T = TA. Moreo ver, such antiferromagne tic spin dynamics can be controlled by \nmagnetic field due to the finite ma gnetic moment at T = TA, which opens a way of studying field-driven \nantiferromagnetic spin dynamics. Furthermore, the fact that field-driven DW speed exhibits a sharp and \nnarrow peak at TA provides a sim ple but accurate method to determine TA, which has not been possible. \nWe also achieve a fast DW speed near room temperature , opening a possibility for ultra -high speed \ndevice operation at room temperature. We expect that s uch findings are also advantageous for current -\ninduced DW motion in ferrimagnets. A low threshold current density , more than one order of magnitude \nsmaller than that of ferromagnets, has already been demonstrated in ferrimagnets32–33. Therefore, by \ntuning TA, one could obtain high-speed an d low power consumed spintronic device s using ferrimagnet s, \nwhich could even be superior to ferromagnetic system s. To conclude, o ur work suggests that revealing \nand tailoring TA, which has not been paid much attention , is crucial for controlling ferrimagne tic \nmagnetisation dynamics, and therefore could be a key for realising ferrimagnet ic spintronics . Method \nFilm preparation and device fabrication. The studied samples are amorphous thin films of \nGd 23Fe67.4Co9.6 of 30 nm thickness which have been deposited by magnetron sputtering. To avoid \noxidation of the GdFeCo layer, 5 nm Si 3N4 were used as buffer and capping layers, respectively. The \nfilms exhibit an out -of-plane magnetic anisotropy . GdFeCo micro strips with a 100 nm -wide Hall bar \nstructure were fabricat ed using electron beam lithography and Ar ion milling process. A negative tone \nelectron beam resist (maN -2403) was used for lithography at a fine resolution (~5 nm). For current \ninjection, Ti(5 nm)/Au(100 nm) electrodes were stacked on the wire. To make an Ohmic contact, the \nSi3N4 capping layer was removed by weak ion milling before electrode deposition. \nExperimental setup for field -driven domain wall motion . A pulse generator (Picosecond 10, 300B) \nwas used to generate a current pulse to create the DW. 100m A and 10ns current pulse is used to create the \nDW. For field -driven DW motion, 1mA dc current (corresponding current density is 7 109 A·m−2) was \nflowed along the wire to generate anomalous Hall voltage, VH. Yokogawa 7651 was used as a current \nsource. The VH at the Hall cross was recorded by the oscilloscope (Textronix 7354) through the 46 dB \ndifferent ial amplifier. Low temperature probe sta tion was used for measuring the DW motion in a wide \nrange of temperature. \nDW detection technique . We used a time -of-flight measurement of DW propagation to obtain a DW \nspeed in a flow regime. The procedure for measuring the DW speed is as follows . First, a large out -of-\nplane magnetic field Bsat = −200 mT is applied to reset the magnetisation . Next, a drive field Bd, in the \nrange of | BP| < |Bd| < |BC|, is applied in the opposite direction. Here, BP is the pinning field of DW motion \nand BC is the coercive field of the sample . Since the Bd is smaller than the BC, the drive field does not \nreverse the magnetisation or create DWs. N ext, a current pulse (100 mA, 1 0 ns) is injected by a pulse \ngenerator to create a DW next to the contact line through current -induced Oersted field . As soon as the \nDW is create d, the Bd pushes the DW because the Bd is larger than the BP. Then the DW propagates along \nthe wire and passes through the Hall cross region . When the DW passes through the Hall cross, the Hall \nvoltage changes abruptly because the magnetisation state of th e Hall cross reverses as a result of the DW passage. This Hall signal change is recorded by the oscilloscope through the 46dB differential amplifier. \nWe refer to this as a ‘signal trace’. Since the detected Hall voltage change includes a large background \nsignal, we subtract the background from the ‘signal trace’ by measuring a ‘reference trace’. The reference \ntrace is obtained in the same manner as the signal trace, except that the saturation field direction is \nreversed ( Bsat = +200 mT). In this reference t race, no DW is nucleated , so that only the electronic noise \ncan be detected in the oscilloscope in the reference trace . To obtain a sufficiently high signal -to-noise \nratio, we averaged the data from 5 repeated measurements . \nAtomic spin model simulation . We adopt the atomistic model simulation s ince the ferrimagnet consists \nof two magnetic components, i.e., RE and TM on an atomic scale . The one -dimensional Hamiltonian of \nferrimagnet is described by ℋ=𝐴𝑠𝑖𝑚∑𝑺𝒊∙𝑺𝒊+𝟏 𝑖 −𝐾𝑠𝑖𝑚∑(𝑺𝒊∙𝒛̂)2𝑖 +𝜅𝑠𝑖𝑚∑(𝑺𝒊∙𝒚̂)2𝑖 , where 𝑺𝒊 is the \nnormalized magnetic moment at lattice site 𝑖. The odd number of 𝑖 represents a site for TM, and the even \nnumber of 𝑖 represents a site for RE. 𝐴𝑠𝑖𝑚,𝐾𝑠𝑖𝑚,𝜅𝑠𝑖𝑚 denote the exchange, easy -axis anisotropy along the \nz axis, and hard -axis anisotropy, respectively. We sol ve the atomistic LLG equation 𝜕𝑺𝒊\n𝜕𝑡=−𝛾𝑖𝑺𝒊×\n𝑯𝒆𝒇𝒇 ,𝒊+𝛼𝑖𝑺𝒊×𝜕𝑺𝒊\n𝜕𝑡, where 𝑯𝒆𝒇𝒇 ,𝒊=−1\n𝑀𝑖𝜕ℋ\n𝜕𝑺𝒊 is the effective field, γi=gi 𝜇𝐵ℏ⁄ is the gyromagnetic \nratio, and 𝑀𝑖 is the magnetic moment for site 𝑖. We use parameters as 𝐴𝑠𝑖𝑚=7.5𝑚𝑒𝑉 ,𝐾𝑠𝑖𝑚=\n0.3𝑚𝑒𝑉,𝜅𝑠𝑖𝑚=−0.8𝜇𝑒𝑉, damping constant 𝛼𝑇𝑀=𝛼𝑅𝐸=0.004, the lattice constant is 0.4 nm, and \nLand é g-factors for each site are gTM=2.2 and gRE=2. References \n1. MacDonald, A. H. & Tsoi, M. Antiferromagnetic metal spintronics. Phil. Tra ns. R. Soc. A 369, 3098 –\n3114 (2011). \n2. Jungwirth, T., Marti, X., Wadley, P. & Wunderlich, J. Antiferromagnetic spintronics . Nat. Nanotech . \n11, 231-241 (2016 ). \n3. Keffer, F. & Kittel, C. Theory of antiferromagnetic resonance. Phys. Rev . 85, 329 –337 (1952). \n4. Gomon ay, O. , Jungwirth, T. & Sinova , J. High Antiferromagnetic Domain Wall Velocity Induced by \nNéel Spin -Orbit Torques . Phys. Rev. Lett. 117, 017202 (2016). \n5. Shiino, T. et al. Antiferromagnetic Domain Wall Motion Driven by Spin -Orbit Torques. Phys. Rev. \nLett. 117, 087203 (2016). \n6. Tveten, E.G., Qaiumzadeh, A., & Brataas, A. Antiferromagnetic Domain Wall Motion Induced by \nSpin Waves . Phys. Rev. Lett . 112, 147204 (2014 ). \n7. Tveten, E.G., Qaiumzadeh, A., Tretiakov, O. A. & Brataas, A. Staggered Dynamics in \nAntiferromagnets by Collective Coordinates. Phys. Rev. Lett . 110, 127208 (2013) . \n8. Evans , R.F.L. et al. Atomistic spin model simulations of magnetic nanomaterials. J. Phys.: \nCondensed Matter , 26, 103202 (2014) . \n9. Yamaguchi, A. et al. Real -space observation of current -driven domain wall motion in submicron \nmagnetic wires. Phys. Rev. Lett . 92, 077205 (2004). \n10. Parkin, S. S. P., Hayashi, M. & Thomas, L. Magnetic domain -wall racetrack memory. Science 320, \n190–194 (2008). \n11. Schryer, N. L. &Wa lker, L. R. The motion of 180 ° domain walls in uniform dc magnetic fields. J. \nAppl. Phys . 45, 5406 -5421 (1974). \n12. Yang, S. -H., Ryu, K. -S. & Parkin, S. Domain -wall velocities of up to 750 m s−1 driven by exchange -\ncoupling torque in synthetic antiferromagnets . Nature Nanotech . 10, 221 –226 (2015) . \n13. Wangness, R. K. Sublattice effects in magnetic resonance. Phys. Rev . 91, 1085 -1091 (1953). 14. Stanciu, C. D. et al. Ultrafast spin dynamics across compensation points in ferrimagnetic GdFeCo: \nThe role of angular momentum compensation. Phys. Rev. B 73, 220402(R) (2006). \n15. Binder, M. et al. Magnetization dynamics of the ferrimagnet CoGd near the compensation of \nmagnetizat ion and angular momentum. Phys. Rev. B 74, 134404 (2006). \n16. Clarke, D. J. , Tretiakov, O. A. , Chern, G. -W., Bazaliy, Ya. B. & Tchernyshyov , O. Dynamics of a \nvortex domain wall in a magnetic nanostrip: Application of the collective -coordinate approach. Phys. \nRev. B 78, 134412 (2008) . \n17. Thiele, A.A. Steady -State Motion of Magnetic Domains . Phys. Rev. Lett . 30, 230 (1973). \n18. Radu, I. et al. Transient ferromagnetic -like state mediating ultrafast reversal of antiferromagnetically \ncoupled spins, Nature , 472, 205 -208 (20 11). \n19. Okuno, T. et al. Temperature dependence of magnetoresistance in GdFeCo/Pt heterostructure, Appl. \nPhys. Express 9, 073001 (2016). \n20. Tanaka, H., Takayama, S. & Fujiwara, T. Electronic -structure calculations for amorphous and \ncrystalline Gd 33Fe67alloys. Phys. Rev. B 46, 7390 -7394 (1992). \n21. Tsuya, N. Microwave resonance in ferrimagnetic substance. Prog. Theoret. Phys . 7, 263 -265 (1952). \n22. Kittel, C. On the Gyromagnetic Ratio and Spectroscopic Splitting Factor of Ferromagnetic \nSubstances. Phys. Rev . 76, 743 (1949 ). \n23. Scott, G. G. Review of Gyromagnetic Ratio Experiments. Rev. Mod. Phys . 34, 102 (1962). \n24. Min, B. I. and Jang, Y. -R. The effect of the spin -orbit interaction on the electronic structure of \nmagnetic materials. J. Phys. Condens. Matter 3, 5131 (1991). \n25. Yoshim ura, Y. et al. Soliton -like magnetic domain wall motion induced by the interfacial \nDzyaloshinskii –Moriya interaction. Nat. Phys . 12, 157 -161 (2016). \n26. Tono, T. et al. Chiral magnetic domain wall in ferrimagnetic GdFeCo wires. Appl. Phys. Express 8, \n073001 (2 015). \n27. Kim, K. -J. et al. Observation of asymmetry in domain wall speed under transverse magnetic field. \nAPL Mater. 4, 032504 (2016). 28. Ono, T. et al. Propagation of a domain wall in a submicrometer magnetic wire. Science 284, 468 –470 \n(1999). \n29. Volkov, V. V. & B okov, V. A. Domain Wall Dynamics in Ferromagnets. Physics of the Solid State \n50, 199 -228 (2008). \n30. Alebrand, S. et al. Light -induced magnetization reversal of high -anisotropy TbCo alloy films. Appl. \nPhys. Lett. 101, 162408 (2012). \n31. Stanciu, C. D. et al. Subpi cosecond Magnetization Reversal across Ferrimagnetic Compensation \nPoints . Phys. Rev. Lett. 99, 217204 (2007). \n32. Ngo, D -T., Ikeda, K. & Awano, H. Direct Observation of Domain Wall Motion Induced by Low -\nCurrent Density in TbFeCo Wires. Appl. Phys. Express 4, 093002 (2011). \n33. Awano, H., Investigation of domain wall motion in RE -TM magnetic wire towards a current driven \nmemory and logic. J. Magn. Magn. Mater . 383, 50-55 (2015). \n34. Jiang, X., Gao., L. Sun J. Z. & Parkin S. S. P. Temperature Dependence of Current -Induce d \nMagnetization Switching in Spin Valves with a Ferrimagnetic CoGd Free Layer. Phys. Rev. Lett. 97, \n217202 (2006). Acknowledgements \nThis work was partly supported by JSPS KAKENHI Grant Numbers 15H05702, 26870300, 26870304, \n26103002, 25220604 , 2604316 Collaborative Research Program of the Institute for Chemical Research, \nKyoto University, and R & D project for ICT Key Technology of MEXT from the Japan Society for the \nPromotion of Science (JSPS). KJK acknowledges support from the KAIST start -up funding. SKK and YT \nacknowledge the support from the Army Research Office under Contract No. 911NF -14-1-0016 . K.-J.L. \nacknowledges support from Creative Materials Discovery Program through the National Research \nFoundation of Korea (NRF -2015M3D1A1070465) . \n \nAuthor contri butions \nK.-J.K., T.M. , and T.O. planed the study. A.T. grew and optimi sed the GdFeCo film. T.T. fabricated the \ndevice and performed the experiment with the guide of K. -J.K.. T.Okuno , W.-S.H., Y.H. , and S.K. helped \nthe experiment . S.-K.K., K.-J.L., and Y.T. provide theory. S.-H.O., G.G., and K. -J.L. performed the \nnumerical simulation. K.-J.K., S.-K.K., K. -J.L., T.M., and T.O analysed the result s. K.-J.K., S.-K.K., K. -\nJ.L., T.M., and T.O. wrote the manuscript. \n \nAdditional Information \nSupplementary Information is available in the online version of the paper. Reprints and permissions \ninformation is available at www.nature.com/reprints . Correspondence and requests for materials should \nbe addressed to K.-J.K, K.-J.L. and T. O. \nCompeting financial interests \nThe authors declare no competing financial interests. \n Figure Legends \nFigure 1| Schematic illustration of Device structure . Schematic illustration of GdFeCo \nmicrowire . The inset shows schematic illustration of two spin sub-lattices below and above the \nmagnetisation compensation temperature, TM. Blue and red arrows indicate Gd and FeCo moments, \nrespectively. \nFigure 2| Identification of magnetisation compensation temperature TM. a, Anomalous Hall \neffect resistance RH as a function of perpendicular magnetic field BZ for several temperatures as denoted \nin the figure. b,. Coercive field, BC, and the magnitude of the Hall resistance change , ΔRH \n Z Z B R B R R H H H\nwith respective to the temperature. The region shaded in red indicates the \nmagnetisation compensation temperature TM \n \nFigure 3| Field -driven domain wall (DW) dynamics across the angular momentum \ncompensation temperature TA. a. DW speed v as a function of driving field Bd for several \ntemperatures as denoted in the figure. Dashed lines are best fits based on \n0 dB B v . b. DW speed v \nas a function of temperature T for several driving fields as denoted in the figure. c. DW mobility μ as a \nfunction of temperature T. The red and blue shaded regions in b and c indicate the magnetisation \ncompensation temperature, TM, and angular momentum compensation temperature, TA, respectively. \n \nFigure 4| Simulation results of ferri magnetic domain wal l (DW) a. DW speed as a function of \nthe out -of-plane field BZ for various indices (see Table 1). Symbols are numerical results whereas solid \nlines are Eq. (1). Inset shows low field regimes, where vertical dotted lines indicate the Walker \nbreakdown fields. b. Computed DW speed as a function of of δs=s1−𝑠2 at various values of BZ. \n \n \n \n \nFig.1 \n \n \n \n \n \n \nFig.2 \n \n \n \n \n \nFig.3 \n \n \n \n \n \n \n \n \nFig.4 \n \n \n \n \n \nTABLE 1. Parameters used in the numerical simulation. \nIndex 1 2 3 4 5 6 7 8 9 \n𝑴𝑭𝒆𝑪𝒐 (𝒌𝑨/𝒎) 1120 1115 1110 1105 1100 1095 1090 1085 1080 \n𝑴𝑮𝒅(𝒌𝑨/𝒎) 1040 1030 1020 1010 1000 990 980 970 960 \n𝜹𝒔(𝟏𝟎−𝟕𝑱∙𝒔 𝒎𝟑⁄ ) -1.24 -0.93 -0.62 -0.31 0 0.31 0.62 0.93 1.24 \n \n " }, { "title": "2008.03062v2.Cavity_magnon_polariton_based_precision_magnetometry.pdf", "content": "Cavity magnon polariton based precision magnetometry\nN. Crescini,1, 2, a)C. Braggio,2, 3G. Carugno,2, 3A. Ortolan,1and G. Ruoso1\n1)INFN-LNL, Viale dell’Università 2, 35020 Legnaro (PD), Italy\n2)Dipartimento di Fisica e Astronomia, Via Marzolo 8, 35131 Padova, Italy\n3)INFN-Sezione di Padova, Via Marzolo 8, 35131 Padova, Italy\n(Dated: 2 October 2020)\nA photon-magnon hybrid system can be realized by coupling the electron spin resonance of a magnetic material to a\nmicrowave cavity mode. The quasiparticles associated with the system dynamics are the cavity magnon polaritons,\nwhich arise from the mixing of strongly coupled magnons and photons. We illustrate how these particles can be used\nto probe the magnetization of a sample with a remarkable sensitivity, devising suitable spin-magnetometers which\nultimately can be used to directly assess oscillating magnetic fields. Specifically, the capability of cavity magnon\npolaritons of converting magnetic excitations to electromagnetic ones, allows for translating to magnetism the quantum-\nlimited sensitivity reached by state-of-the-art microwave detectors. Here we employ hybrid systems composed of\nmicrowave cavities and ferrimagnetic spheres, to experimentally implement two types of novel spin-magnetometers.\nAmong the most studied types of hybrid systems, an im-\nportant role is played by photon-magnon hybrid systems\n(PMHSs)1,2. These yielded remarkable results in the study\nof light-matter interaction3, and in the last decades emerged\nas promising constituents for new quantum technologies as\nwell4–6. PMHSs have different forms, as they are built with\nmiscellaneous building blocks, but the underlying physics is\nsimilar. In a magnetic field B0, a spin can change its quantum\nstate from\u00001=2 to a +1=2 by absorbing a spin-1 boson, like a\nphoton, and vice versa by emitting one. In this sense, a quanta\nof spin excitation with energy ¯hwm=mBB0can be effectively\ndescribed as a quasiparticle, known as magnon, which can\nturn into a photon of the same energy ¯hwc7. This recipro-\ncal conversion is quantified by the interaction strength gcm,\nknown as vacuum Rabi splitting, which is the rate at which\nmagnons are converted into photons and vice versa. When gcm\nis much larger than the damping rates of the magnon gmand\nof the photon gc, the system is in the strong coupling regime,\nand the quasiparticles arising from this mixing are known as\ncavity magnon polaritons (CMPs)8,9.\nPMHSs are widely investigated for advancing quantum\ninformation science. In this field their importance lies in\nbuilding quantum memories10–16, in converting microwaves\nto optical photons17–21, or in quantum sensing, where the de-\ntection of single magnons was recently demonstrated22–24.\nCMP recently found new applications in the field of non-\nHermitian physics25–27, where they already yielded outstand-\ning results28. Exceptional points, spots of the system’s pa-\nrameter space highly sensitive to external stimulations, can\nbe probed with PMHSs29,30, and new configurations may be\ndesigned to access more exotic phenomena and study their\napplications31,32. The potential of hybrid systems was also\nshown in many other applications of quantum physics33–37.\nA distinguished physical realisation of this model can\nbe obtained by hybridising the microwave photons of\na resonant cavity with the magnons of a ferrimagnetic\ninsulator38–42. Such scheme was implemented with multi-\nple purposes, for example to develop new quantum tech-\na)Electronic mail: nicolo.crescini@phd.unipd.itnologies with qubits10,43, or for microwave-to-optical photon\nconversion20,21, making it an established platform of hybrid\nmagnonics.\nIn the devices described in this letter, we employ cop-\nper cavities as a photonic resonator and Yttrium Iron Garnet\n(YIG) spheres as magnetic material (see Fig. 1a). YIG has the\nexceptionally high electron spin density of 2 \u00021028spin=m3\nalready at room temperature, and a linewidth as narrow as\n1 MHz. This latter value is matched to the one of a typical\ncopper cavity and, thanks to the chosen spherical shape, is not\naffected by geometric demagnetization. Being employed in a\nnumber of microwave and rf devices, YIG is among the most\nwell-known ferrites, and hence is readily available. The mag-\nnetic sample is placed inside the cavity, where the rf magnetic\nfield is maximum for the selected cavity mode, and is mag-\nnetised with a static field B0perpendicular to the cavity one.\nIn this way, the Kittel mode of magnetisation couples to the\nmicrowave cavity photons, and the system exhibits the typi-\ncal anticrossing dispersion relation, of which an example is\nshown in Fig. 1b. The coupling strength depends on the work-\ning frequency, on the microwave mode volume, and on the\nnumber of spins involved40, but it is normally large enough to\nlet the photon (magnon) oscillate into magnon (photon) many\ntimes before being dissipated.\nThis feature of CMP to be a mixed state of microwaves and\nspin excitations allows one to extract information on magnons\nby monitoring photons. In the presence of a strong coupling,\nthe signal transduction is efficient, i. e. without signal loss, as\na spin excitation is more likely converted to a photon and de-\ntected, than it is to be dissipated due to the PMHS losses (see\nFig. 1c for a schematic diagram). Amongst other techniques to\nmeasure spin-waves4, the use of CMP is a particularly simple\napproach which exploits the sensitivity of microwave technol-\nogy and transfers it to the detection of magnons. The strong\ncoupling makes the energy stored in a cavity dependent on\nthe one in the material, so an antenna coupled to the elec-\ntromagnetic field of the cavity gives a simple access to the\nfeatures of the spin system44. Nowadays electronics is ex-\ntremely developed, and the detection of electromagnetic radi-\nation has been brought to the standard quantum limit of linear\namplifiers. At microwave frequencies, Josephson Parametric\nAmplifiers (JPA) were demonstrated to be the best devices toarXiv:2008.03062v2 [quant-ph] 30 Sep 20202\nCavity\nMagnetic sample\nMagnetic field\nMagnet(a) (b)\nCavity modeKittel mode\n(c)\nExternal \ninputReadoutPhotons\nMagnonsStrong coupling\nFIG. 1. Schematic representation of a typical PMHS (a), anticrossing\ncurve (b), and diagram of a spin-magnetometer working principle (c).\nPart (a) represents a PMHS consisting in a YIG sphere housed in a\nmicrowave cavity under a static magnetic field. Plot (b) is measured\nwith a .5 mm-diameter YIG sphere in a 14 GHz copper cavity, the\ncolour scale is in logarithmic arbitrary units, where blue to yellow is\nlow to high transmission, and the dashed lines show the uncoupled\ncavity and Kittel modes.\nmeasure tiniest amounts of power45. Thanks to CMPs, such\nprecision can be shifted to a magnetic measurement, as the\nelectromagnetic power in the cavity is highly dependent on\nthe magnetisation of the sample when the coupling strength\nlargely exceeds the system dissipations gcm\u001dgm;gc. It fol-\nlows that, under these conditions, the quantum-limited read-\nout of a JPA can be exploited to detect spin excitations.\nAt microwave frequencies, measuring a sample’s magne-\ntization becomes increasingly difficult because of technolog-\nical limitations and fundamental problems, like for example\nradiation damping46–48. In free space, radiation damping con-\nsists in the magnetic dipole emission of a magnetised sample\nwhich, at GHz frequencies, drastically decreases the coher-\nence time, limiting the experimental sensitivity. This effect is\navoided in PMHSs, as the sample is housed in a resonant cav-\nity which removes the damping by inhibiting the phase space\nof the emission49.\nFor all their characteristics, PMHSs emerge as an outstand-\ning platform for precision magnetic measurement, which are\nof interest for a broad range of applications as well as for ap-\nproaching fundamental physics issues. Hereafter, we describe\ntwo types of spin-magnetometers which can be designed with\nhybrid systems, detail their design and report on their opera-\ntion. We notice that a high occupation number of the modes\npermits to treat them as classical oscillators, which is often the\ncase throughout this work, so we rely on a classical treatment\nof the fields. These devices are originally meant to measure\ntiniest oscillation of a sample’s magnetization, related for ex-\nample to a Dark Matter Axion field49,50, but can be used to\nassess many other physical phenomena.\nTransverse spin-magnetometer (TSM). - Let’s now focus\non a hybrid system like the one of in Fig. 1, where a magne-\ntised YIG sphere is placed in a microwave cavity. If an oscil-\nlating electromagnetic, or pseudo-electromagnetic, field b1isoriented perpendicularly to the static field, its quanta can be\nabsorbed by the hybrid magnetic mode. As the magnetization\nvector Mprecesses over the static field, an excitation lying on\nthe precession plane can resonantly interact with it, and the\nsystem evolves according to Bloch equations\ndM\ndt=g(M\u0002b1)?+M\nTs; (1)\nwhere g= (2p)28GHz =T is the electron gyromagnetic ratio,\nandTsis the system relaxation time. The driven magnetization\nresulting from Eq. (1) is\nM(t) =gmBnsTscos(w1t); (2)\nwhere w1is the frequency of b1. In a steady state, the power of\nb1is absorbed, re-emitted by the magnetization, and rapidly\nconverted into photons thanks to the strong coupling.\nThe optimal experimental condition is an antenna critically\ncoupled to the cavity, which in steady state can extract up to\nhalf of the power deposited by the external field, resulting in\nP1=gmBNsw1b2\n1Ts; (3)\nwhere Nsis the number of spins of the hybrid system, and\nthe field frequency w1is on resonance with one of the hybrid\nmodes. To calculate the magnetic sensitivity of the TSM, in\nEq. (3) we substitute the deposited power P1(in Watts) with\nthe power sensitivity of the readout electronics sP(in Watts\nper unit of bandwidth), and recast the equation to isolate the\nmagnetic field. We obtain the sensitivity of the TSM\nsb1=rsP\ngmBNsw1Ts; (4)\nin Tesla per unit of bandwidth, which is the field detectable\nin 1 second integration time with a unitary signal-to-noise ra-\ntio. Eq. (4) also shows that the spin-magnetometer sensitiv-\nity increases for larger spin-number and longer hybrid system\ncoherence times. This suggests the use of high quality-factor\ncavities and samples to get a long Ts, and of a large volume\nof high spin density magnetic material to increase Ns. In this\nsense, we found a good compromise in YIG. The scalability\nof the PMHS is of fundamental importance to obtain an in-\ncreased sensitivity of the setup, as it is directly related to the\nincrement of Ns. To this aim we design spin-magnetometers\nbased on multi-samples PMHS51,52, embedded in cylindrical\ncavities. To further boost the magnetic sensitivity, we reduce\nsPby operating the device at milli-Kelvin temperatures, to re-\nduce thermal noises and to consent the use of quantum-limited\namplifiers.\nFollowing these directions, we built a TSM whose scheme\nis reported in Fig. 2a. Its PHMS comprises ten YIG spheres,\nall of 2.1 mm-diameter, produced in-house. These are biased\nwith a magnetic field supplied by a superconducting mag-\nnet, with 7 ppm uniformity over the volume containing the\nspheres. We realise the PMHS by placing the spheres along\nthe axis of a cylindrical cavity (33 mm-diameter, 65 mm-\nlength) allowing them to couple with the uniform rf magnetic\nfield of the TM110 mode at 10.7 GHz.3\nThe PMHS has been designed to reduce the effects of the\nmagnetic dipole interaction between different spheres and of\nhigher order magnetostatic modes. By removing the degener-\nacy of the TM110 mode we limit the interference of other cav-\nity modes; this is achieved employing a cavity with a quasi-\ncircular section53,54. To describe this system we used a sec-\nond quantisation model consisting in four coupled harmonic\noscillators. We fit it to the experimental anticrossing curve of\nFig. 2b53. We then operate the magnetometer in the frequency\nband 10.2-10.4 GHz, part of the lower frequency hybrid mode\nrange, as identified by the fit (dashed line in the figure)52. The\noperational range is matched with the working band of our\nJosephson Parametric Converter (JPC), i.e. a JPA formed by\na Josephson ring modulator shunted with four inductances55.\nThe JPC tuning is allowed by a small superconducting coil\nbiased with a constant current, as shown Fig. 2c. The dashed\nlines in Fig. 2c includes the 10.2-10.4 GHz frequency interval,\nshowing that in this range the lower frequency hybrid mode\ncan be monitored with our amplifier. The JPC is screened\nfrom external disturbances with different layers of supercon-\nducting and m-metal shields, and we verified that the solenoid\nproviding the static field is not affecting the resonance fre-\nquencies of the amplifier.\nThe noise temperature and gain of the electronics chain has\nbeen characterised with the injection of microwave signals of\nknown amplitude in an antenna weakly coupled to the cavity.\nThe effective noise temperature results Tn'1K, which sets\nthe noise power per unit of bandwidth sP=kBTn, where kBis\nBoltzmann constant. The contribution of the quantum limit to\nthe noise budget is 0.5 K, and the remaining 0.5 K is consis-\ntent with extra noise added by the second-stage amplifier, by\nthe losses of the wires and by the PMHS thermodynamic tem-\nperature of\u0018100mK52. The spin number and relaxation time\nare obtained by fitting our model to the transmission measure-\nments of Fig. 2b. The measurement of sP, and of NsandTs\nthrough the PMHS spectroscopy, allows us to calculate the\nsensitivity of the TSM using Eq. (4). With the parameters of\nthis setup we obtain a magnetic sensitivity of\nsb1=0:9\u000210\u000018h\u00101K\nTn\u0011\u0010Ns\n1021\u0011\n\u0010w1=2p\n10:4GHz\u0011\u0010Ts\n168ns\u0011i1=2Tp\nHz:(5)\nThat the sensitivity given by Eq. (4) holds if the field to be\ndetected has two characteristics: a coherence time longer than\nTs, and a coherence length long enough to comprise all the Ns\nspins.\nIn particular, this is the case of the field induced by Dark\nMatter axions49,50, which at GHz frequencies satisfies both\nthese conditions. We used this TSM with a fixed band-\nwidth of 5 kHz to search for axions, obtaining a limit on\ntheir effective field of 5 :5\u000210\u000019T with about ten hours of\nintegration52. A TSM has the advantage of being sensitive to\na (pseudo)magnetic field acting on a sample which is within\nthe volume of a resonant cavity. In such a controlled envi-\nronment, external electromagnetic disturbances are unlikely\nto be present, making it an interesting testbed for fundamen-\ntal physics, which are usually not subjected to such screening.\nYIG sphere\nCopper cavity\nSuperconducting magnet\nB-field profile of the \nTM110 modeIsolators JPAReadout\n(b)\n(c)(a)\nHybrid mode\nWorking frequencies[dB]\n[deg]Dipole antenna\nB0 directionFIG. 2. (a) Simplified scheme of the TSM operated for an axion\nsearch52(see text for details). (b) Anticrossing curve of the 10 YIG\nspheres PMHS, where the dashed line indicates the low-frequency\nhybrid mode monitored during the measurement. (c) Phase-current\ndiagram of the JPC mounted in this setup, here the dashed line shows\nthe optimal working points of the amplifier. From plots (b) and (c)\none notes that the 10.2-10.4 GHz band enables both the PMHS signal\ntransduction and the JPA amplification.\nHowever, from the point of view of the TSM possible tech-\nnological employment, this feature is a limitation. In fact, the\nscreening due to the cavity makes it difficult to expose the ma-\nterial to a field which is uniform and coherent over the mag-\nnetic material volume. Hence, the application of this device is\nprobably limited to the search of new physics.\nLongitudinal spin-magnetometer (LSM). - In another pos-\nsible measurement scheme a persistent oscillating B-field is\nparallel to the static one. In this configuration, the sample’s\nmagnetization precesses about a field B0+b2sin(w2t), where\nw2andb2are the oscillating field frequency and amplitude,\nandtis time. To illustrate the experimental arrangement, we\nfirst consider a simplified scheme including only the mate-4\nrial and ignoring the presence of the cavity. The experimen-\ntal scheme is shown in Fig. 3a, where a sphere is surrounded\nby two crossed loops. Loop number 1 is used to excite the\nmaterial, while loop number 2 senses the transmitted rf sig-\nnal, and S21plots are measured. The electron spin resonance\n(ESR) frequency wmof the magnetised sample is modulated\nat the frequency w2\u001cwmby varying the field b2\u001cB0. If\na monochromatic tone is applied on resonance with wm, the\neffect of b2is then to transfer some of the pump power, the\ncarrier, to sidebands at frequencies wm\u0006nw2, as schemati-\ncally shown in Fig. 3a for n=1. In the S21spectrum of this\nsimplified system, the amplitude of the first order sideband\nresults\nz1=pA2\npQb2\n2B0; (6)\nwhere Apis the carrier amplitude and Q=wm=gmthe qual-\nity factor of the ESR. In a standard ESR technique an exter-\nnally applied b2is used to detect the derivative of the ESR\ncurve with a lock-in amplifier. Here we invert such scheme,\nand search for oscillating b2-fields by sensing the presence of\nsidebands. The detection of sidebands is limited by the ef-\nfective noise temperature of the system determining sP, the\npower sensitivity already defined in the case of the TSM. The\namplitude z1is given by Eq. (6) only within the linewidth of\nthe ESR, and drastically reduces for w2>gm. On the other\nhand, when w2 shows a summation over all pairs of the nearest -neighboring sites of different \nsublattices and JAB > 0 and JAC < 0 (model the ferro -ferrimagnetic interactions ) are the nearest -\nneighbor exchange constants. h(t) is the oscillating external magnetic field and is described by \nℎ(𝑡)=ℎ0cos(𝑤𝑡), where h0, w and t are the time, amplitude and angular frequency. 𝜉𝑗 is \ndistributed random variables and it takes the value of unity or zero, according to whether site j is \nfilled by an ion of B or C, respectively. So, 𝜉𝑗 is described by \n \n \n 1 1 , (2) j j jP p pξ ξ ξ \n \n where p and (1−𝑝) are the concentration of B and C ions, respectively. A mixed ferro -\nferrimagnetic ABpC1-p ternary alloy system is in contact with an isothermal heat bath at an absolute \ntemperature Tabs and evolves according to the Glau ber-type stochastic process at a rate of 1/ τ. From \nthe master equation associated to the stochastic process, it follows that the average magnetization \nsatisfies the following equation [39 -45], \n \n \n11 , (3a)2 A A B C\nAB j AC j 0\njjdτ S S tanh β J p S +J S p h cos wtdt\n \n \n\n3\n, (3b)\n21\n \nA\nAB i 0\ni BB\nA\nAB i 0\nisinhβ J p S +ph cos wt\ndτ S Sdtcoshβ J p S +ph cos wt \n \n \n 3 1.5 1 0.5 1, (3c)2 1.5 1 2 0.5 1 CCsinh p βx sinh p βx dτ S Sdt cosh p βx cosh p βx \n \nwhere 𝑥=𝐽𝐴𝐶∑𝑆𝑖𝐴\n𝑖 +ℎ0cos(𝑤𝑡). Using the mean -field theory; the dynamic mean -field \napproximation equations are obtained as follows \n \n \n 11 , (4a)2 A A AB AB A AC AC C 0dm m tanh β J z m p+J z m p h cosd \n \n 3\n, (4b)\n21 AB BA A 0\nBB\nAB BA A 0sinhβ J z m p+ph cos dmmd coshβ J z m p+ph cos\n \n \n 3 1.5 1 0.5 1, (4c)2 1.5 1 2 0.5 1 CCsinh p βy sinh p βy dmmd cosh p βy cosh p βy\n \n \nwhere 𝑦=𝐽𝐴𝐶𝑧𝐶𝐴𝑚𝐴+ℎ0cos(𝜉), 𝜉=𝑤𝑡, Ω=𝜏𝑤 and Ω=2𝜋, zAB, zBA, zAC and zCA are taken \n4 for a square lattice. \n 𝑀𝑖=1\n𝜏∫𝑚𝑖(𝜉)𝑑𝜉 (5) \n \n where 𝑖=𝐴,𝐵 and 𝐶. In other words, 𝑀𝐴, 𝑀𝐵 and 𝑀𝐶 correspond to the dynamic order parameters \nof the magnetic components A, B and C. The total magnetization of the system is \n \n 𝑀𝑇=(𝑀𝐴+𝑀𝐵+𝑀𝐶)\n2 (6) \n \nThe physical parameters have been scaled in terms of 𝐽𝐴𝐵. For example, reduced temperature and \nfield amplitude are respectively defined as 𝑇=𝑘𝐵𝑇𝑎𝑏𝑠\n𝐽𝐴𝐵, and ℎ=ℎ0\n𝐽𝐴𝐵, throughout the \n \n3- Results and Discussion \n \nThe effects of the concentration ratio p and the exchange interaction ratio 𝑅 (|𝐽𝐴𝐶|\n𝐽𝐴𝐵) on dynamic \nmagnetization and DPT of the ternary alloy have been examined. It should be noted that the p = 0 \ncase corresponds to a ferrimagnetic mixed spin -1/2 and spin -3/2 system while for p = 1, \ncorresponds to a mixed spin -1/2 and spin -1 ferromagnetic system. The phase diagram of the \nternary alloy in ( 𝑅−𝑇𝐶) and ( 𝑝−𝑇𝐶) planes are shown in Fig . 2 and Fig . 3, respectively. In these \nfigures, upper graphs are plotted for ℎ=0.1 and the lower ones are plotted for ℎ=0.5. The \ncritical temperature value (phase transition temperature) is a little decrease d with increasing h. \n(𝑇𝐶=2.71 for h=0.1, 𝑇𝐶=2.67 for h=0.5) The r eason for this situation is that , the higher field \namplitude becomes dominant against the ferromagnetic and antiferromagnetic nearest -neighbor \nbonds. It can be seen that from the figures that 𝑇𝐶 increases as 𝑅 increases and the 𝑇𝐶 values do \nnot change with 𝑅 for 𝑝= 1.0. Because the system become an AB alloy, there is no AC \ninteraction. Therefore t he system becomes independent from 𝑅. \n \nIn this section, the effects of p and R on the magnetization of a ternary alloy of the type ABpC1-p \nare discussed. Fig. 4 shows the total magnetization chancing with scaled temperature for R=0.5, \n1.0 and R=2.0 values. It is again seen from the figures that a ll the total magnetization curves merge \nat a unique transition temperature for p = 1.0. For the p =0.0 case, Tc=0 at R =0.0 (see Fig . 1 and \n2) and the dynamic critical temperature of the system increase with an increasing of R. For p=0.25, \nthe antiferromagnetic exchange interaction between the A and C magnetic components becomes \neffective in the system for larger R values . In other words , the 𝐽𝐴𝐶 interaction becomes dominant \nand because 𝐽𝐴𝐶 is negative, the 𝑀𝑇 results are negative. It is noted that the saturation values of \nmagnetization increase for p=0.25 and decrease for p=0.75 with the increasing of R. A second \norder phase transition occurs in the system for R=0.5 and R=1.0. But for R=2, first the system gives \n the first order phase transition and then the second order phase transition occurs. 𝑀𝑇 decreases as \nR increases in the range 0.5 < R <1 at p = 0.5 and after 𝑅>1 (AC interaction is more dominant), \nas R increases, 𝑀𝑇 orientation increases by changing. \n \nIn Fig. 5 the results have been depicted for R=1.0. As expected, t he sign of the 𝑀𝐴 magnetization \nis negative ( 𝑀𝐴=−1/2) 𝑀𝐵=0.0 and 𝑀𝐶=3/2 at T=0 for p=0.0. For 0 .0 0 (FM) \nJBC < 0 (AFM) \nSB=1 \nSC = 3/2 \n \nh=0.1\nR0.0 0.5 1.0 1.5 2.0 2.5 3.0TC\n02468\np=0.00 \np=0.25 \np=0.50 \np=0.75 \np=1.00 \nh=0.5\nR0.0 0.5 1.0 1.5 2.0 2.5 3.0TC\n02468\np=0.00 \np=0.25 \np=0.50 \np=0.75 \np=1.00 \n \nFig. 2: \n \nh=0.1\np0.0 0.2 0.4 0.6 0.8 1.0TC\n02468\nR=0.0 \nR=0.5\nR=1.0\nR=1.5\nR=2.0\nR=2.5\nR=3.0\nh=0.5\np0.0 0.2 0.4 0.6 0.8 1.0TC\n0246R=0.0 \nR=0.5\nR=1.0\nR=1.5\nR=2.0\nR=2.5\nR=3.0\n \n \nFig. 3: \n \nT0 1 2 3 4MT\n0.00.20.40.60.8\np=0.00\np=0.25\np=0.50\np=0.75\np=1.00\nR=JAC/JAB=0.5\nR=JAC/JAB=1.0\nT0 1 2 3 4MT\n-0.4-0.20.00.20.40.60.8\np=0.00\np=0.25\np=0.50\np=0.75\np=1.00\nR=JAC/JAB=2.0\nT0 1 2 3 4MT\n-0.6-0.4-0.20.00.20.40.60.81.0\np=0.00\np=0.25\np=0.50\np=0.75\np=1.00\n \nFig. 4: \n \np=0.00\nT0.0 0.5 1.0 1.5 2.0 2.5 3.0M\n-1.0-0.50.00.51.01.52.0\nMA\nMB\nMC\np=0.25\nT0.0 0.5 1.0 1.5 2.0 2.5 3.0M\n-2.0-1.5-1.0-0.50.00.51.01.5\nMA\nMB\nMC\np=0.50\nT0.0 0.5 1.0 1.5 2.0 2.5 3.0M\n-2.0-1.5-1.0-0.50.00.51.01.5\nMA\nMB\nMC\np=0.75\nT0.0 0.5 1.0 1.5 2.0 2.5 3.0M\n-2.0-1.5-1.0-0.50.00.51.01.5\nMA\nMB\nMC\np=1.00\nT0.0 0.5 1.0 1.5 2.0 2.5 3.0M\n0.00.20.40.60.81.01.2\nMA\nMB\nMC\n \n \nFig. 5: \n \n " }, { "title": "1706.08488v1.Perpendicular_magnetic_anisotropy_in_insulating_ferrimagnetic_gadolinium_iron_garnet_thin_films.pdf", "content": "Perpendicular magnetic anisotropy in insulating ferrimagnetic gadolinium iron garnet\nthin \flms\nH. Maier-Flaig,1, 2S. Gepr ags,1Z. Qiu,3, 4E. Saitoh,3, 4, 5, 6, 7R. Gross,1, 2, 8\nM. Weiler,1, 2H. Huebl,1, 2, 8and S. T. B. Goennenwein1, 2, 8, 9, 10\n1Walther-Mei\u0019ner-Institut, Bayerische Akademie der Wissenschaften, Garching, Germany\n2Physik-Department, Technische Universit at M unchen, Garching, Germany\n3WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan\n4Spin Quantum Recti\fcation Project, ERATO, Japan Science and Technology Agency, Sendai, Japan\n5Institute for Materials Research, Tohoku University, Sendai, Japan\n6PRESTO, Japan Science and Technology Agency, Saitama, Japan\n7Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, Japan\n8Nanosystems Initiative Munich, M unchen, Germany\n9Institut f ur Festk oper- und Materialphysik, Technische Universit at Dresden, Dresden, Germany\n10Center for Transport and Devices of Emergent Materials, Technische Universit at Dresden, 01062 Dresden\n(Dated: June 27, 2017)\nWe present experimental control of the magnetic anisotropy in a gadolinium iron garnet (GdIG)\nthin \flm from in-plane to perpendicular anisotropy by simply changing the sample temperature.\nThe magnetic hysteresis loops obtained by SQUID magnetometry measurements unambiguously\nreveal a change of the magnetically easy axis from out-of-plane to in-plane depending on the sam-\nple temperature. Additionally, we con\frm these \fndings by the use of temperature dependent\nbroadband ferromagnetic resonance spectroscopy (FMR). In order to determine the e\u000bective mag-\nnetization, we utilize the intrinsic advantage of FMR spectroscopy which allows to determine the\nmagnetic anisotropy independent of the paramagnetic substrate, while magnetometry determines\nthe combined magnetic moment from \flm and substrate. This enables us to quantitatively evalu-\nate the anisotropy and the smooth transition from in-plane to perpendicular magnetic anisotropy.\nFurthermore, we derive the temperature dependent g-factor and the Gilbert damping of the GdIG\nthin \flm.\nControlling the magnetization direction of magnetic\nsystems without the need to switch an external static\nmagnetic \feld is a challenge that has seen tremendous\nprogress in the past years. It is of considerable interest\nfor applications as it is a key prerequisite to store infor-\nmation in magnetic media in a fast, reliable and energy\ne\u000ecient way. Two notable approaches to achieve this in\nthin magnetic \flms are switching the magnetization by\nshort laser pulses[1, 2] and switching the magnetization\nvia spin orbit torques[3{5]. For both methods, materials\nwith an easy magnetic anisotropy axis oriented perpen-\ndicular to the \flm plane are of particular interest. While\nall-optical switching requires a magnetization component\nperpendicular to the \flm plane in order to transfer angu-\nlar momentum[2], spin orbit torque switching with per-\npendicularly polarized materials allows fast and reliable\noperation at low current densities[3]. Therefore great ef-\nforts have been undertaken to achieve magnetic thin \flms\nwith perpendicular magnetic anisotropy.[6] However, re-\nsearch has mainly been focused on conducting ferromag-\nnets that are subject to eddy current losses and thus of-\nten feature large magnetization damping. Magnetic gar-\nnets are a class of highly tailorable magnetic insulators\nthat have been under investigation and in use in appli-\ncations for the past six decades.[7{9] The deposition of\ngarnet thin \flms using sputtering, pulsed laser deposition\nor liquid phase epitaxy, and their properties are very well\nunderstood. In particular, doping the parent compound\n(yttrium iron garnet, YIG) with rare earth elements is apowerful means to tune the static and dynamic magnetic\nproperties of these materials.[7, 10, 11]\nHere, we study the magnetic properties of a gadolin-\nium iron garnet thin \flm sample using broadband fer-\nromagnetic resonance (FMR) and SQUID magnetome-\ntry. By changing the temperature, we achieve a transi-\ntion from the typical in-plane magnetic anisotropy (IPA),\ndominated by the magnetic shape anisotropy, to a per-\npendicular magnetic anisotropy (PMA) at about 190 K.\nWe furthermore report the magnetodynamic properties\nof GdIG con\frming and extending previous results.[7]\nI. MATERIAL AND SAMPLE DETAILS\nWe investigate a 2 :6µm thick gadolinium iron gar-\nnet (Gd 3Fe5O3, GdIG) \flm grown by liquid phase epi-\ntaxy (LPE) on a (111)-oriented gadolinium gallium gar-\nnet substrate (GGG). The sample is identical to the\none used in Ref. 12 and is described there in detail.\nGdIG is a compensating ferrimagnet composed of two\ne\u000bective magnetic sublattices: The magnetic sublattice\nof the Gd ions and an e\u000bective sublattice of the two\nstrongly antiferromagnetically coupled Fe sublattices.\nThe magnetization of the coupled Fe sublattices shows a\nweak temperature dependence below room temperature\nand decreases from approximately 190 kA m\u00001at 5 K to\n140 kA m\u00001at 300 K.[8] The Gd sublattice magnetization\nfollows a Brillouin-like function and decreases drasticallyarXiv:1706.08488v1 [cond-mat.mtrl-sci] 26 Jun 20172\nfrom approximately 800 kA m\u00001at 5 K to 120 kA m\u00001at\n300 K.[8] As the Gd and the net Fe sublattice magneti-\nzations are aligned anti-parallel, the remanent magneti-\nzations cancel each other at the so-called compensation\ntemperature Tcomp = 285 K of the material.[13] Hence,\nthe remanent net magnetization Mof GdIG vanishes at\nTcomp.\nThe typical magnetic anisotropies in thin garnet \flms\nare the shape anisotropy and the cubic magnetocrys-\ntalline anisotropy, but also growth induced anisotropies\nand magnetoelastic e\u000bects due to epitaxial strain have\nbeen reported in literature.[14, 15] We \fnd that our ex-\nperimental data can be understood by taking into ac-\ncount only shape anisotropy and an additional anisotropy\n\feld perpendicular to the \flm plane. A full determina-\ntion of the anisotropy contributions is in principle pos-\nsible with FMR. Angle dependent FMR measurements\n(not shown) indicate an anisotropy of cubic symmetry\nwith the easy axis along the crystal [111] direction in\nagreement with literature.[16] The measurements sug-\ngest that the origin of the additional anisotropy \feld\nperpendicular to the \flm plane is the cubic magnetocrys-\ntalline anisotropy. However, the low signal amplitude and\nthe large FMR linewidth towards Tcomp in combination\nwith a small misalignment of the sample, render a com-\nplete, temperature dependent anisotropy analysis impos-\nsible. In the following, we therefore focus only on shape\nanisotropy and the additional out-of-plane anisotropy\n\feld.\nII. SQUID MAGNETOMETRY\nSQUID magnetometry measures the projection of the\nmagnetic moment of a sample on the applied magnetic\n\feld direction. For thin magnetic \flms, however, the\nbackground signal from the comparatively thick sub-\nstrate can be on the order of or even exceed the magnetic\nmomentmof the thin \flm and hereby impede the quan-\ntitative determination of m. Our 2:6µm thick GdIG \flm\nis grown on a 500 µm thick GGG substrate warranting a\ncareful subtraction of the paramagnetic background sig-\nnal of the substrate. In our experiments, H0is applied\nperpendicular to the \flm plane and thus, the projection\nof the net magnetization M=m=Vto the out-of-plane\naxis is recorded as M?. Fig. 1 shows M?of the GdIG \flm\nas function of the externally applied magnetic \feld H0.\nIn the investigated small region of H0, the magnetization\nof the paramagnetic substrate can be approximated by\na linear background that has been subtracted from the\ndata. The two magnetic hysteresis loops shown in Fig. 1\nare typical for low temperatures ( T.170 K) and for\ntemperatures close to Tcomp. The hysteresis loops unam-\nbiguously evidence hard and easy axis behavior, respec-\ntively. Towards low temperatures ( T= 170 K, Fig. 1 (a))\nthe net magnetization M=jMjincreases and hence, the\nanisotropy energy associated with the demagnetization\n\feldHshape =\u0000M?[17] dominates and forces the mag-\n−150 −100 −50 0 50 100 150\nµ0H0(mT)-40-2002040M⊥(kA/m)a\n170 K (1)(2)(3)\nMH0z\n−100 mT 10 mTz\n−150 −100 −50 0 50 100 150\nµ0H0(mT)-4-2024M⊥(kA/m)b\n250 K(2)(3)\n10 mTz − HC+ HC\n−100 mTH0M z\n(1)FIG. 1. Out-of-plane magnetization component M?mea-\nsured by SQUID magnetometry. For di\u000berent temperatures,\nmagnetically hard (170K, (a)) and easy (250K, (b)) axis loops\nare observed. The arrows on the data indicate the sweep di-\nrection ofH0. The insets schematically show the magnetiza-\ntion direction MandH0=H0zwith the \flm normal zat\nthe indicated values of H0.\nnetization to stay in-plane. At these low temperatures,\nthe anisotropy \feld perpendicular to the \flm plane, Hk,\ncaused by the additional anisotropy contribution has a\nconstant, comparatively small magnitude. We therefore\nobserve a hard axis loop in the out-of-plane direction:\nUpon increasing H0from\u0000150 mT to +150 mT, Mcon-\ntinuously rotates from the out-of-plane (oop) direction to\nthe in-plane (ip) direction and back to the oop direction\nagain. The same continuous rotation happens for the op-\nposite sweep direction of H0with very little hysteresis.\nFor temperatures close to Tcomp (T= 250 K, Fig. 1 (b)),\nHshape becomes negligible due to the decreasing Mwhile\nHkincreases as shown below. Hence, the out-of-plane di-\nrection becomes the magnetically easy axis and, in turn,\nan easy-axis hysteresis loop is observed: After applying\na large negative H0[(1) in Fig. 1 (a)] MandH0are\n\frst parallel. Sweeping to a positive H0,M\frst stays\nparallel to the \flm normal and thus M?remains con-\nstant [(2) in Fig. 1 (a)] until it suddenly \rips to being\naligned anti-parallel to the \flm normal at H0>+Hc\n[(3) in Fig. 1 (a)]. These loops clearly demonstrate that\nthe nature of the anisotropy changes from IPA to PMA\non varying temperature.3\n0.0 0.5 1.0\nµ0H0(T)0102030ωres/2π (GHz)110K 190K 240Ka\n0.70 0.75µ0H0(T)\n−40040∆S21×103\n110 K0.30.40.5\nµ0H0(T)−0.50.00.5\n∆S21×103\n240 K\n0 50 100 150 200 250 300\nT(K)0.00.5−µ 0Hi(T)b\nHani=−Meff\nHshape =−M⊥\nHk=M⊥−MeffIPA PMA\nRe ImReIm\nFIG. 2. Broadband FMR spectroscopy data reveiling a smooth transition from in-plane to perpendicular anisotropy. (a)\nFMR resonance frequency plotted against H0taken for three di\u000berent temperatures (symbols) and \ft to Eq. (3) (solid lines).\nFor an IPA, a positive e\u000bective magnetization Me\u000b(positivex-axis intercept) is extracted, whereas Me\u000bis negative for a PMA.\n(inset) Exemplary resonance spectra (symbols) at 14 :5 GHz recorded at 110 K and 240 K as well as the \fts to Eq. (1) used\nto determine !res(solid lines). A complex o\u000bset S0\n21has been subtracted for visual clarity, plotted is \u0001 S21=S21\u0000S0\n21.(b)\nAnisotropy \feld Hani=\u0000Me\u000bas a function of temperature (open squares). Prediction for shape anisotropy Hshape based on\nSQUID magnetometry data (solid line) from Ref. 18. The additional perpendicular anisotropy \feld Hk=M?\u0000Me\u000b(red dots)\nincreases to approximately 0 :18 T at 250 K where its value is essentially identical to Hanidue to the vanishing M?.\nIII. BROADBAND FERROMAGNETIC\nRESONANCE\nIn order to quantify the transition from in-plane to per-\npendicular anisotropy found in the SQUID magnetome-\ntry data, broadband FMR is performed as a function of\ntemperature with the external magnetic \feld H0applied\nalong the \flm normal.[19] For this, H0is swept while\nthe complex microwave transmission S21of a coplanar\nwaveguide loaded with the sample is recorded at vari-\nous \fxed frequencies between 10 GHz and 25 GHz. We\nperform \fts of S21to[20]\nS21(H0)j!=\u0000iZ\u001f(H0) +A+B\u0001H0 (1)\nwith the complex parameters AandBaccounting for a\nlinear \feld-dependent background signal of S21, the com-\nplex FMR amplitude Z, and the Polder susceptibility[21,\n22]\n\u001f(H0) =Me\u000b(H\u0000Me\u000b)\n(H\u0000Me\u000b)2\u0000H2\ne\u000b+i\u0001H\n2(H\u0000Me\u000b):(2)\nHere,\ris the gyromagnetic ratio, He\u000b=!=(\r\u00160), and\n!is the microwave frequency and the e\u000bective magneti-\nzationMe\u000b=Hres\u0000!res=(\r\u00160). From the \ft, the res-\nonance \feld Hresand the full width at half-maximum\n(FWHM) linewidth \u0001 His extracted. Exemplary data\nforS21(data points) and the \fts to Eq. (1) (solid lines)\nat two distinct temperatures are shown in the two in-\nsets of Fig. 2 (a). We obtain excellent agreement of the\n\fts and the data. The insets furthermore show that the\nsignal amplitude is signi\fcantly smaller for T= 240 K\nthan for 110 K. This is expected as the signal amplitude\nis proportional to the net magnetization Mof the sam-\nple which decreases considerably with increasing temper-\nature (cf. Fig. 2 (b)). At the same time, the linewidthdrastically increases as discussed in the following section.\nThese two aspects prevent a reliable analysis of the FMR\nsignal in the temperature region 250 K H k\nindicating that shape anisotropy dominates, and the \flm\nplane is a magnetically easy plane while the oop direction\nis a magnetically hard axis. At 240 K (red curve) Me\u000bis\nnegative and hence, the oop direction is a magnetically\neasy axis. Figure 2 (b) shows the extracted Me\u000b(T).\nAt 190 K, Me\u000bchanges sign. Above this tempera-\nture (marked in red), the oop axis is magnetically easy\n(PMA) and below this temperature (marked in blue),\nthe oop axis is magnetically hard (IPA). The knowledge\nofM?(T) obtained from SQUID measurements allows\nto separate the additional anisotropy \feld HkfromMe\u000b4\n1.61.82.02.2g\na\n10−310−210−1αb\n0 50 100 150 200 250\nT(K)100030005000∆ω0/2π(MHz)c\nFIG. 3. Key parameters characterizing the magnetiza-\ntion dynamics of GdIG as a function of T:(a)g-factor\ng=\r~=\u0016B,(b)Gilbert damping constant \u000band(c)inhomo-\ngeneous linewidth \u0001 !0=(2\u0019)\n(red dots in Fig. 2 (b)). Hk=M?\u0000Me\u000bincreases\nconsiderably for temperatures close to Tcomp while at\nthe same time the contribution of the shape anisotropy,\nHshape =\u0000M?trends to zero. For T'180 K,Hkex-\nceedsHshape which is indicated by the sign change of\nMe\u000b. Above this temperature, we thus observe PMA. We\nuse the magnetization Mdetermined using SQUID mag-\nnetometry from Ref. 18 normalized to the here recorded\nMe\u000bat 10 K in order to quantify Hk. The maximal value\n\u00160Hk= 0:18 T is obtained at 250 K which is the highest\nmeasured temperature due to the decreasing signal-to-\nnoise ratio towards Tcomp.\nWe can furthermore extract the g-factor and damp-\ning parameters from FMR. The evolution of the g-factor\ng=\r~\n\u0016Bwith temperature is shown in Fig. 3 (a). We ob-\nserve a substantial decrease of gtowardsTcomp. This is\nconsistent with reports in literature for bulk GIG and can\nbe explained considering that the g-factors of Gd and Fe\nions are slightly di\u000berent such that the angular momen-\ntum compensation temperature is larger than the mag-\nnetization compensation temperature.[23] The linewidth\n\u0001!=\r\u0001Hcan be separated into a inhomogeneous con-\ntribution \u0001 !0= \u0001!(H0= 0) and a damping contribu-\ntion varying linear with frequency with the slope \u000b:\n\u0001!= 2\u000b\u0001!res+ \u0001!0: (4)\nClose toTcomp = 285 K, the dominant contribution to\nthe linewidth is \u0001 !0which increases by more than an\norder of magnitude from 390 MHz at 10 K to 6350 MHz\nat 250 K [Fig. 3 (c)]. This temperature dependence of the\nlinewidth has been described theoretically by Clogston\net al.[24, 25] in terms of a dipole narrowing of the in-homogeneous broadening and was reported experimen-\ntally before[7, 16]. As opposed to these single frequency\nexperiments, our broadband experiments allow to sepa-\nrate inhomogeneous and intrinsic damping contributions\nto the linewidth. We \fnd that in addition to the in-\nhomogeneous broadening of the line, also the Gilbert-\nlike (linearly frequency dependent) contribution to the\nlinewidth changes signi\fcantly: Upon approaching Tcomp\n[Fig. 3 (b)], the Gilbert damping parameter \u000bincreases\nby an order of magnitude. Note, however, that due to\nthe large linewidth and the small magnetic moment of\nthe \flm, the determination of \u000bhas a relatively large\nuncertainty.[26] A more reliable determination of the\ntemperature evolution of \u000busing a single crystal GdIG\nsample that gives access to the intrinsic bulk damping\nparameters remains an important task.\nIV. CONCLUSIONS\nWe investigate the temperature evolution of the mag-\nnetic anisotropy of a GdIG thin \flm using SQUID mag-\nnetometry as well as broadband ferromagnetic resonance\nspectroscopy. At temperatures far away from the com-\npensation temperature Tcomp, the SQUID magnetome-\ntry reveals hard axis hysteresis loops in the out-of-plane\ndirection due to shape anisotropy dominating the mag-\nnetic con\fguration. In contrast, at temperatures close to\nthe compensation point, we observe easy axis hysteresis\nloops. Broadband ferromagnetic resonance spectroscopy\nreveals a sign change of the e\u000bective magnetization (the\nmagnetic anisotropy \feld) which is in line with the mag-\nnetometry measurements and allows a quantitative anal-\nysis of the anisotropy \felds. We explain the qualitative\nanisotropy modi\fcations as a function of temperature by\nthe fact that the magnetic shape anisotropy contribu-\ntion is reduced considerably close to Tcomp due to the re-\nduced net magnetization, while the additional perpendic-\nular anisotropy \feld increases considerably. We conclude\nthat by changing the temperature the nature of the mag-\nnetic anisotropy can be changed from an in-plane mag-\nnetic anisotropy to a perpendicular magnetic anisotropy.\nThis perpendicular anisotropy close to Tcomp in combi-\nnation with the small magnetization of the material may\nenable optical switching experiments in insulating fer-\nromagnetic garnet materials. Furthermore, we analyze\nthe temperature dependence of the FMR linewidth and\ntheg-factor of the GdIG thin \flm where we \fnd values\ncompatible with bulk GdIG[7, 25]. The linewidth can\nbe separated into a Gilbert-like and an inhomogeneous\ncontribution. We show that in addition to the previously\nreported increase of the inhomogeneous broadening, also\nthe Gilbert-like damping increases signi\fcantly when ap-\nproachingTcomp5\nV. ACKNOWLEDGMENTS\nWe gratefully acknowledge funding via the priority pro-\ngram Spin Caloric Transport (spinCAT), (Projects GO\n944/4 and GR 1132/18), the priority program SPP 1601(HU 1896/2-1) and the collaborative research center SFB\n631 of the Deutsche Forschungsgemeinschaft.\nVI. BIBLIOGRAPHY\n[1] C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk,\nA. Tsukamoto, A. Itoh, and T. Rasing, Physical Review\nLetters 99, 1 (2007).\n[2] C.-H. Lambert, S. Mangin, B. S. D. C. S. Varaprasad,\nY. K. Takahashi, M. Hehn, M. Cinchetti, G. Malinowski,\nK. Hono, Y. Fainman, M. Aeschlimann, and E. E. Fuller-\nton, Science 345, 1337 (2014).\n[3] K. Garello, C. O. Avci, I. M. Miron, M. Baumgartner,\nA. Ghosh, S. Au\u000bret, O. Boulle, G. Gaudin, and P. Gam-\nbardella, Applied Physics Letters 105, 212402 (2014).\n[4] I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten,\nM. V. Costache, S. Au\u000bret, S. Bandiera, B. Rodmacq,\nA. Schuhl, and P. Gambardella, Nature 476, 189 (2011).\n[5] A. Brataas, A. D. Kent, and H. Ohno, Nature Materials\n11, 372 (2012).\n[6] S. Ikeda, K. Miura, H. Yamamoto, K. Mizunuma, H. D.\nGan, M. Endo, S. Kanai, J. Hayakawa, F. Matsukura,\nand H. Ohno, Nature Materials 9, 721 (2010).\n[7] B. Calhoun, J. Overmeyer, and W. Smith, Physical Re-\nview107(1957).\n[8] G. F. Dionne, Journal of Applied Physics 42, 2142\n(1971).\n[9] J. D. Adam, L. E. Davis, G. F. Dionne, E. F. Schloemann,\nand S. N. Stitzer, IEEE Transactions on Microwave The-\nory and Techniques 50, 721 (2002).\n[10] K. P. Belov, L. A. Malevskaya, and V. I. Sokoldv, Soviet\nPhysics JETP 12, 1074 (1961).\n[11] P. R oschmann and W. Tolksdorf, Materials Research\nBulletin 18, 449 (1983).\n[12] H. Maier-Flaig, M. Harder, S. Klingler, Z. Qiu, E. Saitoh,\nM. Weiler, S. Gepr ags, R. Gross, S. T. B. Goennen-\nwein, and H. Huebl, Applied Physics Letters 110, 132401\n(2017).\n[13] G. F. Dionne, Magnetic Oxides (Springer US, Boston,\nMA, 2009).\n[14] S. A. Manuilov, S. I. Khartsev, and A. M. Grishin, Jour-\nnal of Applied Physics 106, 123917 (2009).\n[15] S. A. Manuilov and A. M. Grishin, Journal of Applied\nPhysics 108, 013902 (2010).\n[16] G. P. Rodrigue, H. Meyer, and R. V. Jones, Journal of\nApplied Physics 31, S376 (1960).\n[17] We use the demagnetization factors of a in\fnite thin \flm:\nNx;y;z= (0;0;1).\n[18] S. Gepr ags, A. Kehlberger, F. D. Coletta, Z. Qiu, E.-J.\nGuo, T. Schulz, C. Mix, S. Meyer, A. Kamra, M. Al-\nthammer, H. Huebl, G. Jakob, Y. Ohnuma, H. Adachi,\nJ. Barker, S. Maekawa, G. E. W. Bauer, E. Saitoh,\nR. Gross, S. T. B. Goennenwein, and M. Kl aui, Nature\nCommunications 7, 10452 (2016).\n[19] The alignment of the sample is con\frmed at low temper-\natures by performing rotations of the magnetic \feld di-\nrection at \fxed magnetic \feld magnitude while recording\nthe frequency of resonance !res. As the shape anisotropydominates at low temperatures, !resgoes through an\neasy-to-identify minimum when the sample is aligned\noop.\n[20] H. Maier-Flaig, S. T. B. Goennenwein, R. Ohshima,\nM. Shiraishi, R. Gross, H. Huebl, and M. Weiler, arXiv\npreprint arXiv:1705.05694 .\n[21] J. M. Shaw, H. T. Nembach, and T. J. Silva, Physical\nReview B 87, 054416 (2013).\n[22] H. T. Nembach, T. J. Silva, J. M. Shaw, M. L. Schneider,\nM. J. Carey, S. Maat, and J. R. Childress, Physical\nReview B 84, 054424 (2011).\n[23] R. K. Wangsness, American Journal of Physics 24, 60\n(1956).\n[24] A. M. Clogston, Journal of Applied Physics 29, 334\n(1958).\n[25] S. Geschwind and A. M. Clogston, Physical Review 108,\n49 (1957).\n[26] For the given signal-to-noise ratio and the large\nlinewidth, \u000band \u0001!0are correlated to a non-\nnegligible degree with a correlation coe\u000ecient of\nC(intercept;slope) = \u00000:967." }, { "title": "1903.10271v1.Octahedral_tilting_and_emergence_of_ferrimagnetism_in_cobalt_ruthenium_based_double_perovskites.pdf", "content": "arXiv:1903.10271v1 [cond-mat.mtrl-sci] 25 Mar 2019Octahedral tilting and emergence of ferrimagnetism in coba lt-ruthenium based double\nperovskites\nManjil Das, Prabir Dutta, Saurav Giri and Subham Majumdar∗∗\nSchool of Physical Sciences, Indian Association for the Cul tivation of Science,\n2A & B Raja S. C. Mullick Road, Jadavpur, Kolkata 700 032, INDI A\nRare earth based cobalt-ruthenium double perovskites A 2CoRuO 6(A = La, Pr, Nd and Sm)\nwere synthesized and investigated for their structural and magnetic properties. All the compounds\ncrystallize in the monoclinic P21/nstructure with the indication of antisite disorder between Co\nand Ru sites. While, La compound is already reported to have a n antiferromagnetic state below\n27 K, the Pr, Nd and Sm systems are found to be ferrimagnetic be lowTc= 46, 55 and 78 K\nrespectively. Field dependent magnetization data indicat e prominent hysteresis loop below Tcin\nthe samples containing magnetic rare-earth ions, however m agnetization does not saturate even\nat the highest applied fields. Our structural analysis indic ates strong distortion in the Co-O-Ru\nbond angle, as La3+is replaced by smaller rare-earth ions such as Pr3+, Nd3+and Sm3+. The\nobserved ferrimagnetism is possibly associated with the en hanced antiferromagnetic superexchange\ninteraction in the Co-O-Ru pathway due to bond bending. The P r, Nd and Sm samples also show\nsmall magnetocaloric effect with Nd sample showing highest v alue of magnitude ∼3 Jkg−1K−1at\n50 kOe. The change in entropy below 20 K is found to be positive in the Sm sample as compared\nto the negative value in the Nd counterpart.\nI. INTRODUCTION\nSince the discovery of low field room temperature\nmagneto-resistance in Sr 2FeMoO 61, the double per-\novskites (A 2BB′O6) have been intensely studied2. These\nquaternary compounds can be synthesized with a vary-\ning combinations of cations at the A (alkaline earth or\nrare earth metals) and B/B′(3d, 4dor 5dtransition met-\nals) sites, which provides a scope to access diverse ma-\nterial properties within the similar crystallographic en-\nvironment. Elements with partially filled dlevel at the\nB/B′can give rise to wealth of magnetic ground states\nincluding ferromagnetism, antiferromagnetism, ferrimag-\nnetism as well as glassy magnetic phase. Double per-\novskites are also associated with intriguing electronic\nproperties3such as half-metallic behaviour4, tunneling\nmagneto-resistance5, metal-insulator transition6and so\non.\nIn case of insulating A 2BB′O6double perovskites, the\nmagnetic interaction between B and B′ions is primarily\nB-O-B′superexchange type and Goodenough-Kanamori\nrule can predict the sign of the interaction7. Ideal dou-\nble perovskites have cubic symmetry, but the presence of\ncations with small ionic radii at the A site can distort the\nstructure. Such distortion lowers the lattice symmetry to\ntetragonal or monoclinic, where the BO 6/B′O6octahe-\ndra get tilted through the bending of B-O-B′bond angle.\nIt has been found that for two fixed B and B′, the mag-\nnetic ground state is very much sensitive to this bond\nangle. Doping at the A site can change the bond an-\ngle and henceforth the nature of the ordered magnetic\nstate. For example, substitution of Ca at the Sr site\nof Sr2CoOsO 6drives the system from an antiferromag-\nnetic (AFM) insulator to spin-glass (SG) and eventually\nto a ferrimagnetic (FI) state on full replacement of Sr\nby Ca8,9. There is also report of drastic change in fer-\nrimagnetic coercivity in (Ca,Ba) 2FeReO 6under hydro-static pressure due to the buckling of Fe-O-Re bond10.\nIt has been argued that the magnetic ground state in\nthese insulating systems is an outcome of the compe-\ntition between the interactions along the paths B-O-B′\nand B-O-B′-O-B (and similarly, B′-O-B-O-B′). When\nthe octahedral tilting is minimal, the B-O-B′-O-B type\ninteraction dominates, giving rise to strong AFM corre-\nlations within B sublattice11. However, with increasing\ndistortion, B-O-B′becomes stronger with the simultane-\nous weakening of B-O-B′-O-B coupling, and a simple FI\nstate emerges (provided the magnetic moments at B and\nB′ions are unequal) due to the strong AFM coupling be-\ntween B and B′ions. The intermediate spin-glass state\npossibly arises from these competing interactions.\nRecently, La 2CoRuO 6compound has been shown to\nhave an AFM ground state, which crystallizes in the dis-\ntortedmonoclinicstructure12. Interestingly, the isostruc-\ntural Y 2CoRuO 6shows ferrimagnetism, and turns into a\nspin-glass on La doping at the Y site13. It is therefore\nworthwhile to study the magnetic states of A 2CoRuO 6\nwith different rare-earth atoms at the A site. It is well\nknownthattheionicradiusofrare-earth(inthe3+state)\ndiminishes with increasing atomic number, which can\ntune the lattice distortion. The main goal of the present\nwork is to study the effect of lattice distortion and as-\nsociated change in the magnetic properties with varying\nrare-earth ion at the A site. It is also worthwhile to note\nhow the rare-earth moment is affecting the ground state\nmagnetic properties. In a recent report, the magnetic\nproperties of A 2CoMnO 6compounds were found to get\nstrongly affected by the variation of rare-earth ion at the\nA-site14.2\n/s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48/s32/s73\n/s111/s98/s115\n/s32/s73\n/s99/s97/s108\n/s32/s73\n/s111/s98/s115/s45/s73\n/s99/s97/s108\n/s32/s66/s114/s97/s103/s103/s32/s112/s111/s115/s105/s116/s105/s111/s110/s115\n/s32/s32/s73/s110/s116/s101/s110/s115/s105/s116/s121 /s40/s97/s46/s117/s46/s41\n/s50 /s40/s100/s101/s103/s114/s101/s101/s41/s98\n/s99\n/s97/s91/s97/s93 /s91/s98/s93\n/s32/s73\n/s111/s98/s115\n/s32/s73\n/s99/s97/s108\n/s32/s73\n/s111/s98/s115/s45/s73\n/s99/s97/s108\n/s32/s66/s114/s97/s103/s103/s32/s112/s111/s115/s105/s116/s105/s111/s110/s115\n/s32/s32\n/s50 /s40/s100/s101/s103/s114/s101/s101/s41\nFIG. 1. (a) and (b) show powder x-ray diffraction data of LCRO a nd SCRO respectively collected at room temperature.\nThe inset of (a) shows the crystal structure of the sample. Gr een, pink, blue and spheres indicate La, Ru, Co and O atoms\nrespectively. The bottom panel indicates the octahedral ti lt of four compositions.\nII. EXPERIMENTAL DETAILS\nSingle phase polycrystalline A 2CoRuO 6samples (for\nA = La, Pr, Nd and Sm) were synthesized by solid state\ntechnique. Stiochiometric amounts of A 2O3(except Pr-\nsample, where Pr 6O11was used), Co 3O4and RuO 2were\nwell mixed in a agate morter pestle and calcined at 1073\nK for12h and then sinteredat 1473K for24h in the pel-\nlet form with one intermediate grinding. The powder X-\nray diffraction (PXRD) data were obtained in RIGAKU\nX-ray diffractometer with Cu-K αradiation in the range\n15◦to 80◦. Magnetization ( M) measurements were car-\nried out using a SQUID magnetometer (Quantum De-\nsign, MPMS-3) up to 70 kOe. High field magnetic mea-\nsurements (up to 150 kOe) was performed on a vibrat-\ning sample magnetometer from Cryogenic Ltd., UK. The\ntemperature dependence of the electrical resistivity ( ρ)\nwas measured by DC four-probe method in the temper-\nature range between 50 K and 300 K.III. SAMPLE CHARACTERIZATION\nAll four compositions, La 2CoRuO 6(LCRO),\nPr2CoRuO 6(LCRO) Nd 2CoRuO 6(NCRO) and\nSm2CoRuO 6(SCRO), crystallize in monoclinic rock salt\nstructure (space group P21/n)15. For double perovskite\nA2BB′O6, one can define a Goldschmidt tolerance factor\nt=rA+rO √\n2(rB+rO), whererA, andrOare the ionic radii of\nA and O respectively, while rBstands for the average\nradius of B and B′. It has been found that if t <1,\nthere can be distortion from the ideal cubic structure2.\nFor the present case, r3+\nLa= 103.2 pm, r3+\nPr= 99 pm r3+\nNd=\n98.3 pm, r3+\nSm= 95.8 pm, r2+\nCo= 65(74.5) pm for low-spin\n(high-spin), r4+\nRu= 62 pm and r2−\nO= 140 pm, which give\ntin the range of 0.81-0.83 for four samples. Evidently\nfor these compositions, tis significantly lower than unity,\nand this explains the observed monoclinic symmetry\nrather than ideal cubic one. This lower symmetry is\nassociated with the tilting of the (B,B′)O6octahedra.\nIn order to determine the crystallographic parameters\nof our samples, we have performed Reitveld refinement\non the room temperature PXRD data using MAUD\nsoftware package16. The data converges well with3\n/s48 /s56/s48 /s49/s54/s48 /s50/s52/s48 /s51/s50/s48/s48/s46/s48/s50/s48/s46/s48/s52\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s46/s48/s49/s50/s48/s46/s48/s49/s56/s48/s46/s48/s50/s52\n/s48 /s56/s48 /s49/s54/s48 /s50/s52/s48 /s51/s50/s48/s48/s56/s49/s54\n/s48 /s56/s48 /s49/s54/s48 /s50/s52/s48 /s51/s50/s48/s48/s49/s54/s51/s50\n/s50/s52 /s52/s56 /s55/s50/s48/s50/s48/s52/s48\n/s48 /s50/s48 /s52/s48 /s54/s48/s48/s50/s52/s48 /s56/s48 /s49/s54/s48 /s50/s52/s48 /s51/s50/s48/s48/s56/s49/s54\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s46/s48/s50/s53/s48/s46/s48/s53/s48/s40/s101/s109/s117/s47/s109/s111/s108/s41/s40/s101/s109/s117/s47/s109/s111/s108/s41/s32/s49/s48/s48/s32/s79/s101\n/s32/s90/s70/s67\n/s32/s70/s67/s32/s90/s70/s67\n/s32/s70/s67\n/s32\n/s84 /s32/s40/s75/s41/s32/s32/s40/s101/s109/s117/s47/s109/s111/s108/s41\n/s91/s97/s93/s76/s67/s82/s79\n/s84/s32 /s40/s75/s41/s32/s49/s48/s48/s32/s79/s101\n/s32/s32\n/s84/s32 /s40/s75/s41\n/s78/s67/s82/s79\n/s32/s49/s48/s48/s32/s79/s101/s84/s32 /s40/s75/s41/s40/s101/s109/s117/s47/s109/s111/s108/s41/s32\n/s32\n/s32/s32\n/s91/s99/s93/s84\n/s80 /s84\n/s80/s83/s67/s82/s79\n/s32/s49/s48/s48/s32/s79/s101/s32/s90/s70/s67\n/s32/s70/s67/s32\n/s32\n/s32/s32\n/s91/s100/s93\n/s32/s32/s32/s40/s101/s109/s117/s47/s109/s111/s108/s41\n/s84/s32 /s40/s75/s41/s32/s49/s48/s48/s32/s79/s101\n/s32/s53/s48/s48/s32/s79/s101\n/s32/s49/s48/s48/s48/s32/s79/s101\n/s32/s32/s32\n/s32/s32/s49/s32/s107/s79/s101/s84\n/s80\n/s40/s101/s109/s117/s47/s109/s111/s108/s41\n/s32/s49/s48/s48/s32/s79/s101/s32/s90/s70/s67\n/s32/s70/s67/s80/s67/s82/s79\n/s91/s98/s93\n/s32/s32\n/s32/s49/s48/s48/s32/s79/s101\n/s32/s32\nFIG. 2. (a)-(d) represent temperature variation of magneti c susceptibility for LCRO, PCRO, NCRO and SCRO respectively\nfor an applied field of 100 Oe both in ZFC and FC protocols. The i nsets of (a) and (b) show the modified Curie-Weiss fit to\nthe susceptibility data of respective samples. The inset of (c) shows the susceptibility of NCRO measured under 1 kOe of fi eld.\nThe inset in (d) shows an enlarged view of the ZFC susceptibil ity of SCRO recorded at different applied fields.\nmonoclinic space group P2 1/n for all the samples along\nwith antisite disorder between Co and Ru sites [Figs. 1\n(a) and (b)]. Our calculations show that there are 1-5%\nantisite defects, i.e., the fractional occupancy of B site\nconsists of 99-95% Co and 1-5% Ru. The antisite defect\nis found to be large in Nd and Sm compounds and it\nis low in case of La and Pr counterparts. The refined\ncrystallographic parameters are depicted in Table 1,\nand they match well with the previous reported data17.\nWe have also added the structural data of Y 2CoRuO 6\n(YCRO) from reference13for comparison. The crystal\nstructure of LCRO, as obtained from the refinement of\nour PXRD data, is shown in the inset of fig. 1 (a). It is\nclearly evident that the (Co,Ru)O 6octahedra are tilted.\nThe average tilting angle is defined as /angbracketleftΨ/angbracketright=1\n2[π−/angbracketleftΦ/angbracketright],\nwhere/angbracketleftΦ/angbracketrightis the average inter-octahedral Co-O-Ru\nangle. Clearly, the tilt angle increases systematically aswe movefrom La to Sm, which is the effect of the gradual\nbending of Co-O-Ru bond (see Table 1). It is interesting\nto note that all the crystallographic parameters vary\nsystematically with the ionic radius for La, Pr, Nd and\nSm samples.\nIV. RESULTS\nIV.1. Magnetic studies\nFigs. 2 (a)-(d) depict the temperature ( T) dependence\nof susceptibility ( χ=M/H) measured under different\nvalues of Hvalues for all the four samples, where both\nzero-field-cooled (ZFC) and field-cooled (FC) measure-\nments were performed. LCRO [fig.2 (a)] shows well de-\nfined peak at the Ne´ el temperature TN= 27 K, indicat-4\nParameters LCRO PCRO NCRO SCRO YCRO\nr3+\nA(˚A) 1.032 0.990 0.983 0.958 0.900\na(˚A) 5.575(2) 5.497(1) 5.440(5) 5.390(2) 5.266\nb(˚A) 5.638(7) 5.689(7) 5.717(7) 5.722(8) 5.711\nc(˚A) 7.886(2) 7.802(4) 7.739(2) 7.685(7) 7.558\nβ(◦) 90.01(1) 89.88(7) 89.99(8) 89.93(2) 90.03\n/angbracketleft∠Co-O-Ru /angbracketright(◦)152.4 150.9 143.4 139.5 141.7\n/angbracketleftΨ/angbracketright(◦) 13.8 14.6 18.3 20.3 19.7\nMag. state AFM FI FI FI FI\nTran. temp. TN= 27 K Tc= 46 K Tc= 55 K Tc= 78 K Tc= 82 K\nHcoer –9 kOe (2 K) 8 kOe (2 K) 22 kOe (2 K) 22.5 kOe (5 K)\nTABLE I. Crystallographic lattice parameters ( a,b,c, andβ), average Co-O-Ru bond angle, average octahedral tilt angl e (Ψ),\nmagnetic state, magnetic transition temperatures and coer civity are depicted for A 2CoRuO 6(A = La, Pr, Nd, Sm and Y). The\nparameters for Y compound are obtained from reference13. The ionic radii of rare-earth ions (in the 3+ state with coor dination\nnumber VI) at the A site are also shown18.\ning an AFM ground state and it matches well with the\nprevious report7,12,19. TheTvariation of susceptibility\n(χ=M/H) of LCRO in the paramagnetic (PM) state\ncannot be fitted with a simple Curie-Weiss law. How-\never, the χ(T) data above 100 K can be fitted well with\na modified Curie-Weiss law, χCW(T) =C/(T−θ)+χ0,\nwherean additional Tindependent term ( χ0) is included.\nHereCis the Curie constant and θis the Curie-Weiss\ntemperature. The effective PM moment, µeff, obtained\nfrom Curie-Weiss fitting, is found to be 6.63 µB/f.u. The\nvalue ofθis−140 K, signifying strong AFM correlations.\nThe FC and ZFC data show weak divergence below TN.\nWe also observed an upward rise in the χ(T) data below\n6 K. The observed value of µeffis higher than expected\nfor a Co2+-high-spin and Ru4+-low-spin states7, which\nwas attributed to extended 4 d-orbitals of Ru19.\nFor PCRO, NCRO and SCRO, the χ(T) data are dras-\ntically different from that of LCRO [figs. 2 (b), (c) and\n(d) respectively], and it isquite eventful. χshowsasharp\nrise below Tc= 46, 55 and 78 K for these magnetic rare-\nearth containing samples respectively. The FC and ZFC\nsusceptibilities show strong irreversibility below Tc. The\ndivergence exists even in measurement at H= 1 kOe [see\ninset of fig. 2 (c)], however the extend of divergence re-\nduces. The point of bifurcation also moves to lower T\nwith increasing H. The ZFC data show a well defined\npeak at a temperature TP, which lies below Tc. Notably,\nwe observe a change in the value of TPwith increasing\nH. TheHdependence of TPis particularly significant\nfor SCRO, where we observe a shift of Tpby 16 K when\nHis changed from 100 Oe to 1 kOe [see inset of fig. 2\n(c)]. Similar shift in the observed peak in ferrimagnetic\nNd2CoMnO 6was also reported previously, which was at-\ntributed to the presenceofferromagnetic(FM) and AFM\nclusters20. TheFCsusceptibility, ontheotherhand, rises\nmonotonically for all the samples with decreasing T.\nThe susceptibility data of PCRO and NCRO can be\nwell fitted with χCW(T) above 100 K, which gives the\nvalues of µeffto be 6.59 and 6.47 µBrespectively. Thevalue of θis -25 (-22) K for Pr(Nd) sample. These val-\nues ofµeffare slightly lower than the expected value of\n6.97 (7.01) µBwith Pr3+(Nd3+), Co2+(high-spin) and\nRu4+(low-spin) states. This mismatch can be caused\nby the presence of antisite disorder. Such disorder is ex-\npected to cause a decrease in the magnetic moment, and\nempirically Mactual=Mobs/(1−2D)21whereDis the\ndegree of disorder between Co and Ru atoms. For exam-\nple, in case of NCRO with D= 0.05 and Mobs= 6.47\nµB,Mactualis found to be 7.18 µB, which is very close\nto the theoretically predicted value.\nFor SCRO, we failed to achieve good fit using χCW(T)\nin the temperature range 100-315 K. The separation be-\ntween ground ( J= 5/2) and first excited ( J= 7/2) mul-\ntiplets in Sm3+is small, and their mixing can be respon-\nsible for the observed non-Curie-Weiss behaviour22.\nThe isothermal MversusHdata are shown in figs. 3\n(a)to(d). ForLCRO,alinear M−Hcurveisobtainedat\n2 K [see fig. 3 (a)], indicating AFM state. On the other\nhand, Pr, Nd and Sm compounds show significant hys-\nteresis with large coercive field ( Hcoer). Non-zero Hcoer\nis observed for these FI systems just below Tc, and it in-\ncreases with decreasing T. The values of Hcoerare found\nto be 9, 8 and 22 kOe at 2 K for PCRO, NCRO and\nSCRO respectively. However, Mdoes not saturate even\nat 70 kOe of field for none the samples. We have also\nrecorded high field magnetization for SCRO, as shown in\nthe inset of fig. 3 (d). The M−Hcurve at 5 K does\nnot fully saturate even at 150 kOe of applied field. Mat-\ntains a value close to 2 µBat the highest H. The value of\nHcoerfor SCRO at 5 K is found to be 10.6 kOe, which is\nsmaller than the value of Hcoerreported for Y 2CoRuO 6\n(∼22.5 kOe) at the same T.\nConsideringnon-zerocoercivityandthepresenceofan-\ntisite disorder,wehaverecorded M−Hhysteresisloopat\n2 K after the sample being field-cooled from room tem-\nperature. In case of inhomogeneous magnetic systems, a\nshift in the hysteresis loop along the field axis may be\nobserved due to the interfacial coupling of two magnetic5\n/s45/s55/s48 /s45/s51/s53 /s48 /s51/s53 /s55/s48/s45/s51/s48/s51/s45/s55/s48 /s45/s51/s53 /s48 /s51/s53 /s55/s48/s45/s48/s46/s52/s48/s46/s48/s48/s46/s52\n/s45/s55/s48 /s45/s51/s53 /s48 /s51/s53 /s55/s48/s45/s49/s46/s54/s48/s46/s48/s49/s46/s54\n/s45/s49/s53/s48 /s45/s55/s53 /s48 /s55/s53 /s49/s53/s48/s45/s50/s48/s50/s45/s55/s48 /s45/s51/s53 /s48 /s51/s53 /s55/s48/s45/s51/s48/s51/s77 /s32/s40\n/s66/s47/s102/s46/s117/s46 /s41\n/s52/s53/s32/s75/s50/s48/s32/s75/s50/s32/s75\n/s32\n/s32/s32/s77 /s32/s40\n/s66/s47/s102/s46/s117/s46 /s41\n/s91/s99/s93/s78/s67/s82/s79/s91/s97/s93/s32/s50/s32/s75\n/s32\n/s32/s32\n/s76/s67/s82/s79\n/s32/s83/s67/s82/s79/s50/s32/s75/s32\n/s91/s100/s93\n/s72 /s40/s107/s79/s101/s41\n/s32/s32\n/s53/s32/s75\n/s32/s32 /s32/s32/s80/s67/s82/s79/s52/s48/s32/s75/s49/s48/s32/s75/s50/s32/s75\n/s91/s98/s93/s32 /s32/s32\nFIG. 3. (a) to (d) show isothermal magnetization data up to fie ld 70 kOe at different temperatures for LCRO, PCRO, NCRO\nand SCRO respectively. The inset of (d) shows the M−Hcurve for SCRO at 5 K for maximum field of 150 kOe.\nphases, and it is referred as exchange bias effect23. Many\ndouble perovskites show exchange bias effect due to the\npresence of antisite disorder24. However, we failed to ob-\nserve such exchange bias in NCRO and SCRO samples,\nwhich possibly rule out the existence of large magnetic\ninhomogeneity in the system.\nIn orderto investigatethe effect ofexternal field on the\nmagnetic state, we have measured magneto-caloric effect\n(MCE) ofthe samples in terms ofentropy-change(∆ SM)\nbyH. In the recent past, MCE has emerged out to be\nan important technique for green refrigeration25. In the\npresent work, we have obtained MCE from our isother-\nmal magnetization data recorded at different constant\ntemperatures. From the theory of thermodynamics,\n∆SM(0→H0) =/integraldisplayH0\n0/parenleftbigg∂M\n∂T/parenrightbigg\nHdH,\nwhere ∆SM(0→H0) denotes the entropy change for thechange in Hfrom 0 to H026. Figs. 4 (a) to (c) show\n∆SM(T,H0) versus Tplot at different values of H0for\nPCRO, NCRO and SCRO samples respectively.\nThe magnitude of MCE is found to be low for all three\nsamples. For the Pr and Nd samples, ∆ SM(T) is mostly\nnegative with its magnitude peaking around 12 and 8\nK (peak magnitude: 1.7 and 2.9 Jkg−1K−1atH0=\n50 kOe) respectively. A broad feature is also observed\nin ∆SM(T,H0) data around 35 K. On the other hand,\nSCRO shows a contrasting behaviour as far as the MCE\nis concerned. ∆ SM(T) for SCRO is positive below 20 K,\nand increases with decreasing temperature (at least for\nH0>10 kOe). ∆ SMattains a value of 3.1 Jkg−1K−1for\nH0= 50 kOe at 2 K.\nAs already discussed, double perovskite systems can\nshow glassy magnetic state27. An intermediate SG state\nis observed when the AFM state is transformed into an\nFI state by suitable A-site doping. In order to investi-6\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s45/s51/s45/s50/s45/s49/s48/s49\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48/s45/s49/s48/s49/s50/s51\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48 /s53/s48/s45/s50/s45/s49/s48/s49\n/s91/s98/s93\n/s32/s32\n/s84 /s32/s40/s75/s41/s32/s50/s48/s32/s107/s79/s101\n/s32/s51/s48/s32/s107/s79/s101\n/s32/s52/s48/s32/s107/s79/s101\n/s32/s53/s48/s32/s107/s79/s101\n/s91/s97/s93/s78/s67/s82/s79\n/s91/s99/s93/s32\n/s32/s83/s67/s82/s79\n/s32/s32\n/s32/s49/s48/s32/s107/s79/s101\n/s32/s50/s48/s32/s107/s79/s101\n/s32/s51/s48/s32/s107/s79/s101\n/s32/s53/s48/s32/s107/s79/s101/s32 /s32/s83\n/s77/s32/s40/s74/s107/s103/s45/s49\n/s75/s45/s49\n/s41\n/s32/s32/s50/s48/s32/s107/s79/s101\n/s32/s51/s48/s32/s107/s79/s101\n/s32/s52/s48/s32/s107/s79/s101\n/s32/s53/s48/s32/s107/s79/s101/s80/s67/s82/s79\n/s32/s32\nFIG. 4. (a) to (c) respectively show the Tvariation of ∆ SMof PCRO, NCRO and SCRO at different magnetic fields.\ngate the possibility of a glassy state, particularly those\nwhich lie acrossthe AFM-FI boundary, field-cooled-field-\nstopmemorymeasurementwereperformedon Prand Nd\nsamples28,29. In this protocol, the samples were cooled\nin 100 Oe of field down to 2 K with intermediate stops at\nseveral temperatures below Tc. Subsequently, the sam-\nples were heated back in 100 Oe and dc magnetization\nwas measured. We do not observe any anomaly at the\nstopping temperatures during heating, which rules out\nthe possibility of any glassy (spin glass or cluster glass)\nor super paramagnetic state in PCRO and NCRO sam-\nples.\nIV.2. Electrical transport\nLikemanyotherA 2BB′O6compounds, LCRO,PCRO,\nNCRO and SCRO show semiconducting behaviour as ev-\nident from the transport data depicted in figs. 5 (a) to\n(d) respectively. The values of ρat room temperature\n(∼300 K) are found to be 8.86, 8.81, 4.16 and 6.68 Ω-\ncm for La, Pr, Nd and Sm compounds. Our analysis on\ntheρ(T) data indicates that all three compositions show\nMott Variable Range (VRH) hopping conduction, where\nρ(T)∼exp/bracketleftBig/parenleftbigT0\nT/parenrightbig1\n4/bracketrightBig\n. This is also quite common among\ndisordered double perovskite such as Sr 2MnRuO 630or\nSr2CoSbO 631. The VRH type conduction is quite clearly\nvisible from the log ρversusT−1/4plots in the respec-\ntive insets of figs. 5 (a), (b) and (c). While for La and\nNd compounds, the VRH nature is present almost over\nthe full range of temperature (it deviates only below 65\nK), Pr and Sm compounds show VRH conduction only\nin the range 160 to 60 K. The values of the parameter T0\nassociated with the VRH conduction are found to be 6.9\n×107, 5.1×107, 8.8×107and 6.8×107K for La, Pr, Nd\nand Sm compounds respectively.V. DISCUSSIONS\nIt turns out that the magnetic properties of samples\ncontaining magnetic rare-earth are drastically different\nfrom that of LCRO. Naively, one can relate the FI state\nwith the magnetic moment of A site. However, FI state\nis also observed in Y 2CoRuO 6, where A site contains\nnonmagnetic Y3+ions. In order to address the issue, let\nus first discuss various salient observations made on the\nstudied samples.\n1. The Co-O-Ru bond distortion (and consequently,\nthe octahedral tilt) is found to increase as we move\nfrom LCRO to heavier rare-earth containing com-\npounds. This is due to the reduction of ionic radius\nat the A-site (lanthanide contraction), as we pro-\nceed from La to Sm. It has been alreadymentioned\nthat the smaller radius of A-element leads to larger\nB-O-B′bond distortion2.\n2. LCRO with relatively smaller bond distortion\nshows AFM ground state. On the other hand,\nPCRO, NCRO and SCRO show large increase in\nMbelow the magnetic transition at Tc. Isothermal\nmagnetization curves show large hysteresis with\nthe presenceofsignificant remanentmagnetization.\nIsostructural Y 2CoRuO 6shows similar magnetic\nbehaviour and the Co-O-Ru superexchange inter-\naction is found to be AFM in nature leading to a\nFI ground state (Co and Ru sublattices have differ-\nent moment values)13. In analogy with the Y com-\npound, we can conclude that the ground states of\nPCRO, NCRO and SCRO are ferrimagnetic. Sim-\nilar to Y compound, FI state in PCRO and subse-\nquent compounds emerges possibly due to the Co-\nO-Ru bond distortion, rather than A-site magnetic\nmoment formation.\n3. All four studied compounds are found to be semi-\nconductingwithreasonablyhighresistivity( ∼MΩ-\ncm)aroundthemagneticanomalies. Therefore,the\nmagnetic interaction is likely to be mediated by the7\n/s55/s53 /s49/s53/s48 /s50/s50/s53 /s51/s48/s48/s48/s46/s48/s48/s48/s46/s53/s50/s49/s46/s48/s52\n/s48/s46/s50/s53 /s48/s46/s51/s48 /s48/s46/s51/s53/s52/s56/s49/s50/s55/s53 /s49/s53/s48 /s50/s50/s53 /s51/s48/s48/s48/s46/s48/s48/s48/s46/s51/s51/s48/s46/s54/s54\n/s48/s46/s50/s52 /s48/s46/s51/s48 /s48/s46/s51/s54/s52/s56/s49/s50\n/s55/s53 /s49/s53/s48 /s50/s50/s53 /s51/s48/s48/s48/s46/s48/s48/s46/s50/s48/s46/s52\n/s48/s46/s50/s53 /s48/s46/s51/s48 /s48/s46/s51/s53/s52/s56/s49/s50/s55/s53 /s49/s53/s48 /s50/s50/s53 /s51/s48/s48/s48/s46/s48/s48/s48/s46/s50/s56/s48/s46/s53/s54\n/s48/s46/s50/s52 /s48/s46/s51/s48 /s48/s46/s51/s54/s52/s56/s49/s50\n/s83/s67/s82/s79/s78/s67/s82/s79/s91/s99/s93/s32\n/s32/s32/s32/s108/s110/s32\n/s84 /s32/s45/s49/s47/s52\n/s32/s40/s75/s32/s45/s49/s47/s52\n/s41 /s32/s76/s67/s82/s79/s40 /s45/s99/s109/s41\n/s40 /s41/s40 /s45/s99/s109/s41\n/s84 /s32/s45/s49/s47/s52\n/s32/s40/s75/s32/s45/s49/s47/s52\n/s41 /s32/s108/s110/s32\n/s32/s32\n/s91/s97/s93\n/s32\n/s32\n/s32/s32\n/s91/s100/s93\n/s108/s110/s32\n/s32/s32\n/s32/s32\n/s84 /s32/s45/s49/s47/s52\n/s32/s40/s75/s32/s45/s49/s47/s52\n/s41 /s32\n/s32/s32/s80/s67/s82/s79/s91/s98/s93\n/s84 /s32/s45/s49/s47/s52\n/s32/s40/s75/s32/s45/s49/s47/s52\n/s41 /s32/s108/s110/s32\n/s32/s32\n/s32/s32\nFIG. 5. (a) to (d) respectively show resistivity data as a fun ction of temperature for LCRO, PCRO, NCRO and SCRO. The\ninsets show the VRH-type fittings to the data.\nsuperexchange, rather than the double-exchange\nmechanism.\n4. The coercivity associated with M−Hhysteresis\nloop is found to be much higher in case of SCRO,\nhowever it is lower than the value reported for\nY2CoRuO 6. It appears that the coercivity is not\ndirectly connected to the fact whether A site con-\ntains a magnetic (such as Pr, Nd or Sm) or non-\nmagnetic (here Y) ion.\n5. ∆SMversusTplots for PCRO, NCRO and SCRO\nshow further anomaly at around 8-12 K. It is diffi-\ncult to guess the origin of such anomaly. However,\nconsideringthe presence of rare-earthat the A-site,\nit may signify the low- Tordering of the magnetic\nrare-earth ions.\nAs already mentioned, a transition from AFM to FI\nvia a glassy magnetic state has been observed in sev-\neral double perovskites with the bending of the B-O-\nB′bond8–10,13, where the lattice distortion was cre-\nated by systematic doping at the A site or by ap-\nplying hydrostatic pressure. In the present case, wefound similar effect when one rare-earth ion is replaced\nby another one with smaller ionic radius. In analogy\nwith the idea mooted in case of Sr 2−xCaxFeOsO 6and\nSr2−xCaxCoOsO 68,9, the AFM state in LCRO is due to\nthe strong AFM correlation along the long bonds Co-O-\nRu-O-Co and Ru-O-Co-O-Ru when the Co-O-Ru bond\ndistortion is low. Replacement of La by Pr initiates\nstrong bending in Co-O-Ru bond (see Table 1), which\npossibly strengthen the Co-O-Ru superexchangeover the\nmagnetic interaction on longer Co-O-Ru-O-Co and Ru-\nO-Co-O-Ru pathways leading to FI state. The bending\nfurther enhances in Sm compound, and a significantly\nlarge coercive field is observed.\nThe above argument is also supported by the drastic\nchange in the values of paramagnetic Curie temperature\nθinPrandNdcompounds(-25and-22Krespectively)as\ncompared to the antiferromagnetically ordered La coun-\nterpart. This may be an indication of the weakening\nof the exchange interaction along longer Co-O-Ru-O-Co\nand Ru-O-Co-O-Ru pathways. However, θstill remains\nnegative due to the presence of Co-O-Ru AFM interac-\ntion.\nIn case of SrCaCoOsO 6and La 2−xYxCoRuO 6(x≈8\n0.25 to 1.5) SG states are observed8,13, when Ca(Y) is\ndoped at the Sr(La) site, which has been assigned due to\nthe frustration between long (B-O-B′-O-B) and short(B-\nO-B′) exchangepaths. However, wedo notsee anyglassy\nmagnetic state in neither of the Pr and Nd compounds.\nFor La 1.25Y0.75CoRuO 6with tilt angle 15.5◦, a promi-\nnent frequency dispersion is observed in the ac suscep-\ntibility data. On the other hand PCRO with lower tilt\nangle (14.6◦) shows ordered FI state. Possibly, the emer-\ngence of glassy state in doped samples is also connected\nwith the doping induced disorder. It is to be noted that\nLa1.25Y0.75CoRuO 6sample has huge antisite disorder of\n20%, as compared to 1% antisite disorder in PCRO.\nIt is now pertinent to address the role of 4 fmoment\nfrom the rare-earth present at the A site. In case of\nisostructural Er 2CoMnO 6, rare-earth moment orders at\na relatively lower Tthan the Co-Mn ordering tempera-\nture32. From our magnetization data, it is hard to iden-\ntify the ordering of rare-earth moment. The ∆ SMversus\nTdata depicted in fig.4 show peak like anomalies be-\ntween 8 and 12 K in PCRO, NCRO and SCRO , and\nthey can be probable ordering points of rare-earth mo-\nments. In order to shed more light on this issue, we have\ncompared the moments of A 2CoRuO 6(A = Pr, Nd, Sm\nand Y), where the magnetic data of Y 2CoRuO 6is ob-\ntained from reference13. The moments at 5 K ( M5K), on\napplying 50 kOe of field, is found to be 2.3, 2.9, 1.4 and\n0.8µB/f.u. on the virgin line of the M−Hcurve for Pr,\nNd, Sm and Y compounds. Y does not carry any mo-ment, while the total angular momentum J= 4, 9/2 and\n5/2 for Pr3+, Nd3+and Sm3+states respectively. The\nmagnetic moments of these three ions are 3.58, 3.62 and\n0.84µBrespectively. Clearly, the variation of M5Kcor-\nresponds well with the variation of rare-earth moment.\nThis signifies that the A site rare-earth moment remains\nin ordered state at 5 K. It is important to perform a\nneutron diffraction study to ascertain the true magnetic\nstructure of these compounds.\nIn summary we have studied the structural, magnetic\nas well as transport properties of the double perovskites\nLa2CoRuO 6, Pr2CoRuO 6, Nd2CoRuO 6, Sm2CoRuO 6.\nWe observe a systematic change in the magnetic ground\nstate as La is replaced by Pr, Nd and Sm. This matches\nwell with the case of Fe-Os and Co-Os based double per-\novskites,wherelatticedistortiontunesthestrengthofthe\nmagnetic interactions in different exchange pathways.\nVI. ACKNOWLEDGMENT\nThe work is supported by the financial grant from\nDST-SERB project (EMR/2017/001058). MD would\nlike to thank CSIR, India for her research fellowship,\nwhile PD thanks DST-SERB for his NPDF fellowship\n(PDF/2017/001061).\nREFERENCES\n∗ ∗sspsm2@iacs.res.in\n1Kobayashi K -I, Kimura T, Sawada H, Terakura K and\nTokura Y 1989 Nature395677\n2Vasala S and Karppinen 2015 Prog. Solid State Chem. 43\n1\n3Nag A, Jana S, Middey S and Ray S 2017 Ind. J. Phys. 91\n883\n4Serrate D, Teresa J M De and Ibarra M R 2007 J. Phys.:\nCondens. Matter 19023201\n5SarmaDD,RaySugata, TanakaK,KobayashiM,Fujimori\nA, Sanyal P, Krishnamurthy H R and Dasgupta C 2007\nPhys. Rev. Lett. 98157205\n6Kato H, Okuda T , Okimoto Y, Tomioka Y, Oikawa K,\nKamiyama T and Tokura Y 2002 Phys. Rev. B 65144404\n7Dass R I, Yan J -Q and Goodenough J B 2004 Phy. Rev.\nB69094416\n8Morrow R, Yan Jiaqiang, McGuire Michael A, Freeland\nJohn W, Haskel Daniel and Woodward Patrick M 2015\nPhys. Rev. B 92094435\n9Morrow R, Freeland J W and Woodward P M 2014 Inorg.\nChem.537983\n10Escanhoela C A Jr., Fabbris G, Sun F, Park C, Gopalakr-\nishnan J, Ramesha K, Granado E, Souza-Neto N M, Vee-\nnendaal M. vanand Haskel D 2018 Phys. Rev. B 98054402\n11Kanungo S, Yan B, Jansen M and Felser C 2014 Phys. Rev.\nB89214414\n12Bos J W G and Attfield J P 2005 J. Mater. Chem. 15715\n13Deng Z et al. 2018 Chem. Mater. 30704714Sahoo RC, Das Sand Nath TK 2018 J. Mag. Mag. Mater.\n460409\n15Anderson M T, Greenwood K B, Taylor G A and Poep-\npelmeiert K R 1993 Prog. Solid St. Chem. 22197\n16http://maud.radiographema.eu\n17Kawano T, Takahashi J, Yamada T and Yamane H 2007\nJ. Ceramic Society of Japan 115792\n18http://abulafia.mt.ic.ac.uk/shannon/ptable.php\n19Yoshii K, Ikeda N and Mizumaki M 2006 phys. stat. sol.\n(a)2032812\n20Das R R, Lekshmi P N, Das S C and Santhosh P N 2019\nJ. Alloys and Compounds 773770\n21Feng H L, Arai M, Matsushita Y, Tsujimoto Y, Guo Y,\nSathish C I, Wang X, Yuan Y -H, Tanaka M and Yamaura\nK 2014J. Am. Chem. Soc. 1363326\n22de Wijn H W, van Diepen A M and Buschow K H J 1973\nPhys. Rev. B 7525\n23Giri Set al2011J. Phys.: Condens. Matter 23073201\n24Coutrim L T, Bittar E M, Baggio-Saitovitch E and Bu-\nfaical L 2017 J. Mag. Mag. Mater. 441243\n25Balli M, Jandl S, Fournier P and Gospodinov M M 2014\nAppl. Phy. Lett. 104232402\n26Pecharsky V K and Gschneidner Jr. K A 1999 J. Mag.\nMag. Mater. 20044\n27Kumar P Anil, Mathieu R, Vijayaraghavan R, Majumdar\nS, Karis O, Nordblad P, Sanyal B, Eriksson O and Sarma\nD D 2012 Phy. Rev. B 860944219\n28Sun Y, Salamon M B, Garnier K, and Averback R S 2003\nPhys. Rev. Lett. 91167206\n29Pramanick S, Chattopadhyay S, Giri S, Majumdar S, and\nChatterjee S 2014 J. Appl. Phys. 116083910\n30Woodward P M et al 2008 J. Am. Ceram. Soc. 91179631Martin V P, Jansen M 2001 J. Sol. Stat. Chem. 15776\n32Blasco J, Subas G, Garca J, Stankiewicz J, Rodr´ ıguez Ve-\nlamaz´ an J A, Ritter C and Garc´ ıa-Mu˜ noz J L 2017 Solid\nState Phenom. 25795" }, { "title": "2007.06805v2.Three_dimensional_Ising_Ferrimagnetism_of_Cr_Fe_Cr_trimers_in_FeCr2Te4.pdf", "content": "arXiv:2007.06805v2 [cond-mat.str-el] 18 Aug 2020Three-dimensional Ising Ferrimagnetism of Cr-Fe-Cr trime rs in FeCr 2Te4\nYu Liu,1R. J. Koch,1Zhixiang Hu,1,2Niraj Aryal,1Eli Stavitski,3Xiao\nTong,4Klaus Attenkofer,3E. S. Bozin,1Weiguo Yin,1and C. Petrovic1,2\n1Condensed Matter Physics and Materials Science Department ,\nBrookhaven National Laboratory, Upton, New York 11973, USA\n2Department of Physics and Astronomy, Stony Brook Universit y, Stony Brook, New York 11790, USA\n3National Synchrotron Light Source II, Brookhaven National Laboratory, Upton, New York 11973, USA\n4Center for Functional Nanomaterials, Brookhaven National Laboratory, Upton, New York 11973, USA\n(Dated: August 19, 2020)\nWe carried out a comprehensive study of magnetic critical be havior in single crystals of ternary\nchalcogenide FeCr 2Te4that undergoes a ferrimagnetic transition below Tc∼123 K. Detailed critical\nbehavior analysis and scaled magnetic entropy change indic ate a second-order ferrimagentic transi-\ntion. Critical exponents β= 0.30(1) with Tc= 122.4(5) K,γ= 1.22(1) with Tc= 122.8(1) K, and\nδ= 4.24(2) at Tc∼123 K suggest that the spins approach three-dimensional Isi ng (β= 0.325, γ\n= 1.24, and δ= 4.82) model coupled with the attractive long-range intera ctions between spins that\ndecay as J(r)≈r−4.88. Our results suggest that the ferrimagnetism in FeCr 2Te4is due to itinerant\nferromagnetism among the antiferromagnetically coupled C r-Fe-Cr trimers.\nINTRODUCTION\nTernaryACr 2X4(A =transitionmetal, X=S, Se, and\nTe) exhibit a variety of magnetic and electronic proper-\nties. The family includes metallic CuCr 2X4and semi-\nconducting Hg(Cd)Cr 2Se4ferromagnets [1–3], semicon-\nducting Fe(Mn)Cr 2S4ferrimagnets [4–6], and insulating\nZnCr2S4antiferromagnet [7]. The FeCr 2X4compounds\nshow competing spin-orbit and exchange interactions [8].\nFeCr2S4is ferrimagnetic (FIM) insulator below Tc= 165\nK, shows a crossover transition from insulator to metal\nnearTcand colossal magnetoresistance behavior [9–11].\nFeCr2Se4is an insulating antiferromagnet (AFM) with\nTN= 218 K and ferrimagnetic with a small magnetic\nmoment of 0.007 µBbelow 75 K [12–14]. It should be\nnoted that FeCr 2Se4crystallizes in the Cr 3S4-type mon-\noclinic structure described within C2/mspace group, in\ncontrast to the cubic spinel-type of FeCr 2S4. However,\nFeCr2S4and FeCr 2Se4have similar electronic structure\nwith nearly trivalent Cr3+and divalent Fe2+states [15].\nThe magnetic moments of Cr ions are antiparallel to\nthoseofFeionsinFeCr 2X4, andthereisstronghybridiza-\ntion between Fe 3 d-states and X p-states [15].\nFeCr2Te4has not been studied much presumably due\nto the difficulty in sample preparation [16–18]. Demeaux\net al. first grew the single crystals of FeCr 2Te4[16]. The\ncrystal structure was reported as a defective NiAs-type\nwithinP63/mmcspace group, where Fe and Cr occupy\nthe same site with alloying ratio of 1 : 2 and a net oc-\ncupancy of 0.75 [16]. In contrast, Valiev et al. reported\nthat FeCr 2Te4crystalizes in a CoMo 2S4-type structure\nwithinI2/mspace group, in which Fe and Cr are octa-\nhedrally coordinated by six Te [17]. Recently, a series of\npolycrystals FeCr 2Se4−xTexwere synthesized [18]. Sub-\nstitution with Te gradually suppresses the AFM order of\nFeCr2Se4, and leads to a short range ferromagnetic(FM)\ncluster metallic state in polycrystal FeCr 2Te4[18]. In or-der to study the intrinsic physical property, high quality\nsingle crystal is required.\nIn this work, we successfully fabricated single crystals\nof FeCr 2Te4and performed a comprehensive study of the\nstructural and magnetic properties. Our analysis of crit-\nicality around Tcindicates that FeCr 2Te4displays the\n3D-Ising behavior, with the magnetic exchange distance\ndecaying as J(r)≈r−4.88. Our first principles calcula-\ntions suggest that the ferrimagnetism in FeCr 2Te4stems\nfrom the itinerant ferromagnetism among the antiferro-\nmagnetically coupled Cr-Fe-Cr trimers. Since transition-\nmetal chalcogenides represent model systems for explor-\ning local structure-related relationship between the bro-\nken symmetry and d-orbital magnetism [19], detailed lo-\ncal stucture investigation of this system would be highly\ndesirable and would bring important new insights.\nEXPERIMENTAL DETAILS\nSingle crystals of FeCr 2Te4were fabricated by melting\nstoichiometric mixture of Fe (99.99%, Alfa Aesar) pow-\nder, Cr (99.95%, Alfa Aesar) powder, and Te (99.9999%,\nAlfa Aesar) pieces. The starting materials were vacuum-\nsealed in a quartz tube, heated to 1200◦C over 12\nh, slowly cooled to 900◦C at a slow rate of 1◦C/h,\nand then quenched into iced water. The single crystal\nx-ray diffraction (XRD) data were taken with Cu Kα\n(λ= 0.15418 nm) radiation of a Rigaku Miniflex pow-\nder diffractometer. In order to obtain more comprehen-\nsive crystallographic information, the powder XRD mea-\nsurements were performed at the PDF beamline (28-ID-\n1) at National Synchrotron Light Source II (NSLS II)\nat Brookhaven National Laboratory (BNL) using Perkin\nElmer image plate detector. The setup utilized x-ray\nbeam with a wavelength of 0.1666 ˚A, and the sample to\ndetectordistance of1.25m, ascalibrated usinga Ni stan-2\ndard. Sample was cooled with an Oxford Cryosystems\n700 cryostream using liquid nitrogen. Raw data were\nintegrated and converted to intensity vs. scattering an-\ngle using the software pyFAI [20]. The average structure\nwas assessed from raw powder diffraction data using the\nGeneralStructureAnalysisSystemII(GSAS-II) software\npackage[21]. Theelementalanalysiswasperformedusing\nenergy-dispersive x-ray spectroscopy (EDS) in a JEOL\nLSM-6500 scanning electron microscope (SEM). The x-\nray absorption spectroscopy (XAS) measurements were\nperformed at 8-ID beamline of the NSLS II (BNL) in a\nfluorescencemode. Thex-rayabsorptionnearedgestruc-\nture (XANES) and extended x-ray absorption fine struc-\nture (EXAFS) spectra were processed using the Athena\nsoftware package. The AUTOBK code was used to nor-\nmalize the absorption coefficient, and separate the EX-\nAFS signal, χ(k), from the atom-absorption background.\nThe extracted EXAFS signal, χ(k), wasweighed by k2to\nemphasize the high-energy oscillation and then Fourier-\ntransformed in krange from 2 to 10 ˚A−1to analyze\nthe data in Rspace. The x-ray photoelectron spec-\ntroscopy (XPS) experiment was carried out in an ultra-\nhigh vacuum system with base pressures <5×10−9Torr,\nequipped with a hemispherical electron energy analyzer\n(SPECS, PHOIBOS 100) and a twin anode x-ray source\n(SPECS, XR50). Al Kα(1486 eV) radiation was used at\n13 kV and 30 mA. The angle between the analyzer and\nthe x-ray source was 45◦and photoelectrons were col-\nlectedalongthesamplesurfacenormal. TheXPSspectra\nwereanalyzedanddeconvolutedusingtheCasaXPSsoft-\nware. The dc/ac magnetic susceptibility were measured\nin Quantum Design MPMS-XL5 system. The applied\nfield (Ha) was corrected as H=Ha−NM, whereMis\nthe measured magnetization and Nis the demagnetiza-\ntion factor. The corrected Hwas used for the analysis of\nmagnetic entropy change and critical behavior.\nRESULTS AND DISCUSSIONS\nIn the single-crystal XRD pattern (inset to Fig. 1),\nonly (00 l) peaks were observed. Synchrotron powder\nXRD pattern of pulverized crystal of FeCr 2Te4can be\nwell fitted by using a monoclinic structure with the I2/m\nspace group (Fig. 1), confirming main phase of FeCr 2Te4\nwith less than 4% FeTe impurity. The determined lattice\nparameters at 300 K are a= 6.822(2)˚A,b= 3.938(1)˚A,\nc= 11.983(5)˚A, andβ= 90.00(5)◦, close to the reported\nvalues [17, 18]. No structural transition was observed on\ncooling, showing a compression along caxis with c=\n11.867(5) ˚A and a slight expansion in the abplane with\na= 6.825(1)˚A andb= 3.943(1)˚A down to 105 K (Table\nI).\nThe ratio of elements in the single crystal as deter-\nmined by EDS is Fe : Cr : Te = 0.99(2) : 1.90(2) : 4.0(1)\n[Fig. 2(a)], and it is referred to as FeCr 2Te4through-\n/s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48\n/s50 /s113 /s32/s40/s100/s101/s103/s46/s41/s40/s48/s48/s50/s41\n/s40/s48/s48/s52/s41/s73/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s46/s32/s117/s46/s41\nFIG. 1. (Color online) Synchrotron powder x-ray diffraction\n(XRD) data and structural model refinements. The data are\nshown by (+) and ( ◦), structural model fit at 300 K and\n105 K is shown by red and blue solid line, respectively. The\ndifference curves are given by green solid lines, offset for cl ar-\nity. The vertical tick marks represent Bragg reflections of t he\nI2/mspace group andupto4% of FeTe impurity. Insetshows\nsingle crystal XRDpattern of FeCr 2Te4at room temperature.\nout this paper. The information on the valence states of\nFe, Cr and Te atoms can be obtained from the element\ncore-level XPS spectra [Fig. 2(b)]: Fe2+(2p3/2∼710\neV, 2p1/2∼723 eV), Cr3+(2p3/2∼575 eV, 2 p1/2∼586\neV), Te2−(3d5/2∼572eV, 3 d3/2∼583 eV)]. Figure 2(c)\nshows the normalized Cr and Fe K-edge XANES spectra,\nin which a similar prepeak feature is observed, indicating\nsimilar local atomic environment for Fe and Cr atoms.\nThe prepeak feature for Fe K-edge is somewhat weaker\nthan that of Cr K-edge, suggesting a weaker lattice dis-\ntortion in FeTe 6when compared with CrTe 6. The edge\nfeatures are close to the standard compounds with Cr3+\nand Fe2+oxidation states [22, 23], in line with the XPS\nresult.\nThe local environment of Fe and Cr atoms is revealed\nin the EXAFS spectra of FeCr 2Te4measured at room\ntemperature [Figs. 2(d) and 2(e)]. In a single-scattering\napproximation, the EXAFS can be described by [24]:\nχ(k) =/summationdisplay\niNiS2\n0\nkR2\nifi(k,Ri)e−2Ri\nλe−2k2σ2\nisin[2kRi+δi(k)],\nwhereNiis the number of neighbouring atoms at a dis-\ntanceRifromthephotoabsorbingatom. S2\n0isthepassive\nelectrons reduction factor, fi(k,Ri) is the backscattering\namplitude, λis the photoelectron mean free path, δiis\nthe phase shift, and σ2\niis the correlated Debye-Waller\nfactor measuring the mean square relative displacement\nof the photoabsorber-backscatter pairs. The corrected3\n/s53/s46/s57/s56 /s54/s46/s48/s48 /s55/s46/s49/s48 /s55/s46/s49/s50 /s55/s46/s49/s52/s48/s46/s48/s48/s46/s53/s49/s46/s48\n/s53/s54/s48 /s53/s56/s48 /s55/s48/s48 /s55/s50/s48\n/s50 /s51 /s52/s48/s49/s50\n/s50 /s51 /s52/s48/s46/s48/s48/s46/s54/s49/s46/s50\n/s50 /s51 /s52 /s53 /s54 /s55 /s56 /s57 /s49/s48/s45/s49/s48/s49\n/s50 /s52 /s54 /s56 /s49/s48/s45/s49/s48/s49\n/s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48\n/s40/s97/s41\n/s78/s111/s114/s109/s97/s108/s105/s122/s101/s100/s32 /s109 /s40/s69/s41\n/s80/s104/s111/s116/s111/s110/s32/s101/s110/s101/s114/s103/s121/s32/s40/s107/s101/s86/s41/s67/s114/s32/s75/s45/s101/s100/s103/s101/s40/s99/s41\n/s70/s101/s40/s98/s41\n/s40/s100/s41 /s40/s101/s41/s51/s48\n/s50/s48\n/s49/s48\n/s48/s73/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s46/s117/s46/s41\n/s66/s105/s110/s100/s105/s110/s103/s32/s69/s110/s101/s114/s103/s121/s32/s40/s101/s86/s41\n/s70/s101/s32/s50/s112\n/s51/s47/s50\n/s70/s101/s32/s50/s112\n/s49/s47/s50/s84/s101/s32/s51/s100\n/s53/s47/s50\n/s84/s101/s32/s51/s100\n/s51/s47/s50/s67/s114/s32/s50/s112\n/s51/s47/s50\n/s67/s114/s32/s50/s112\n/s49/s47/s50\n/s231/s99 /s40/s82/s41 /s231 /s32/s40/s97/s46/s117/s46/s41\n/s82/s40/s197/s41/s40/s103/s41\n/s67/s114/s32/s45/s32/s84/s101/s50/s46/s55/s48/s32/s197/s231/s99 /s40/s82/s41 /s231 /s32/s40/s97/s46/s117/s46/s41/s107/s101/s86/s49/s48 /s49/s50 /s49/s52/s49/s54 /s49/s56/s50/s48/s56/s54/s52 /s50/s73/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s46/s32/s117/s46/s41/s53/s48/s54/s48\n/s52/s48\n/s70/s101 /s67/s114/s84/s101/s70/s101\n/s67/s114/s84/s101\n/s82/s40/s197/s41/s40/s102/s41\n/s70/s101/s32/s45/s32/s84/s101/s50/s46/s53/s56/s32/s197\n/s107/s50\n/s99 /s40/s107/s41/s32/s40/s197/s45/s50\n/s41\n/s107/s32/s40/s197/s45/s49\n/s41\n/s107/s50\n/s99 /s40/s107/s41/s32/s40/s197/s45/s50\n/s41\n/s107/s32/s40/s197/s45/s49\n/s41\nFIG. 2. (Color online) (a) Crystal structure and results of\nEDS analysis on FeCr 2Te4. Room-temperature Fe-2p, Cr-2p,\nand Te-3d core-level XPS (b) and normalized Cr and Fe K-\nedge XANES spectra (c). Fe and Cr K-edge EXAFS oscilla-\ntions (d,e) and Fourier transform magnitudes (f,g) of EXAFS\ndata measured at room temperature. The experimental data\nare shown as blue symbols alongside the model fit plotted as\nred line. Corresponding first coordination shell of Fe and Cr\nare shown in the insets.\nmain peak in the Fourier transform magnitudes of Fe K-\nedge EXAFS around R∼2.58˚A is clearly smaller than\nthat of Cr K-edge EXAFS at R∼2.70˚A [Figs. 2(f) and\n2(g)]. The different local Fe-Te and Cr-Te bond lengths\nsuggest that the Fe and Cr atoms might occupy different\ncrystallographic sites, ruling out the possibility of NiAs-\ntype structure with the same Fe/Cr sites. Then we focus\non the first nearest neighbors of Fe and Cr atoms rang-\ning from 1.5 to 3.5 ˚A. The main peak corresponds to two\nFe-Te bond lengths of 2.69 ˚A and 2.75 ˚A, and four dif-\nferent Cr-Te bond lengths with 2.58 ˚A, 2.67˚A, and 2.80\n˚A, 2.86 ˚A, respectively, extracted from the model fits\nwith fixed coordination number CN. The peaks above\n3.25˚A are due to longer Fe-Cr, Fe-Te, and Cr-Te bond\ndistances, and the multiple scattering involving different\nnear neighbours of the Fe/Cr atoms.\nFigure3(a)showsthe temperaturedependenceofmag-TABLE I. Average and local structural parameters extracted\nfrom the powder XRD and the EXAFS spectra of FeCr 2Te4.\nCN is coordination number based on crystallographic value,\nR is interatomic distance, and σ2is Debye Waller factor.\n300 K 105 K\na(˚A) 6.822(2) 6.825(1)\nb(˚A) 3.938(1) 3.943(1)\nc(˚A) 11.983(5) 11.867(5)\nβ(◦) 90.00(5) 90.01(15)\natom site x y z\nFe 2 a 0 0 0\nCr 4 i 0.001(3) 0 0.2541(8)\nTe1 4 i 0.328(4) 0 0.3726(4)\nTe2 4 i 0.339(4) 0 0.8785(4)\nbond CN R ( ˚A) ∆R ( ˚A) σ2(˚A2)\nCr-Te1 2 2.58 0.01 0.001\nCr-Te2 1 2.67 0.01 0.001\nCr-Te3 1 2.80 0.11 0.003\nCr-Te4 2 2.86 0.11 0.003\nFe-Te1 4 2.69 0.36 0.02\nFe-Te2 2 2.75 0.24 0.02\nnetization measured in out-of-plane field µ0H= 0.1 T,\nin which χincreases with decreasing temperature and in-\ncreases abruptly near Tcdue to the paramagnetic (PM)-\nFIM transition. The in-plane χ(T) [inset in Fig. 3(a)]\nis much smaller than that in out-of-plane field, indicat-\ning the presence of large magnetic anisotropy with easy\ncaxis. The average susceptibility χave= (2/3)χab+\n(1/3)χcfrom 150 to 300 K can be fitted by the Curie-\nWeiss law χave=χ0+C/(T−θ) [Fig. 3(c)], which yields\nχ0= 0.87(1) emu/mol-Oe, C= 7.91(6) emu-K/mol-Oe\n[µeff= 7.95(9)µBper formulaunit], and θ= 126.6(2)K.\nThe positive value of θindicates dominance of ferromag-\nnetic or ferrimagnetic exchangeinteractions in FeCr 2Te4.\nFor monoclinic FeCr 2Se4with a similar layered struc-\nture, the individual spins of Fe and Cr ions have AFM\ncoupling along the caxis with the distance of 2.956 ˚A,\nwhile FM coupling along the baxis with the distance of\n3.617˚A [25]. At low temperature, the FM interaction\ndominating over the AFM interaction results in a FIM\nground state. For FeCr 2Te4, the enhanced hybridization\nbetween d-orbital of Fe and Cr with p-orbital of Te plays\nan important role in magnetic coupling. Based on our\nfirst-principle calculation (see below), the FIM structure\nwhere Fe and Cr atoms have opposite spin orientations\nare the most stable state in FeCr 2Te4, when compared\nwith the FM and AFM structures, which needs further\nverificationby neutron scatteringexperiment. The bifur-\ncation between ZFC and FC curves [Fig. 3(a)] might be\ndue to strong magnetic anisotropy and/or multidomain\nstructure, which has also been observed in other long-4\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s52/s56/s49/s50/s49/s54/s50/s48\n/s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s46/s48/s48/s46/s52/s48/s46/s56/s49/s46/s50/s49/s46/s54/s50/s46/s48\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s53/s49/s48/s49/s53/s50/s48\n/s45/s53/s46/s48 /s45/s50/s46/s53 /s48/s46/s48 /s50/s46/s53 /s53/s46/s48/s45/s52/s45/s50/s48/s50/s52/s32/s90/s70/s67\n/s32/s70/s67/s32/s40/s101/s109/s117/s47/s109/s111/s108/s45/s79/s101/s41\n/s84/s32/s40/s75/s41/s48/s72/s47/s47/s99/s32/s97/s116/s32/s48/s46/s49/s32/s84/s40/s97/s41\n/s48/s72/s47/s47/s97/s98/s32/s97/s116/s32/s48/s46/s49/s84/s32/s90/s70/s67\n/s32/s70/s67/s32/s40/s101/s109/s117/s47/s109/s111/s108/s45/s79/s101/s41\n/s84/s32/s40/s75/s41\n/s97/s118/s101/s32/s40/s109/s111/s108/s45/s79/s101/s47/s101/s109/s117/s41\n/s84/s32/s40/s75/s41/s40/s98/s41\n/s32/s61/s32/s48/s46/s49/s32/s84/s97/s118/s101/s32/s61/s32 /s32/s67/s47/s40/s84/s45 /s41/s109/s39/s32/s40/s101/s109/s117/s47/s109/s111/s108/s41\n/s84/s32/s40/s75/s41/s84\n/s99/s32/s61/s32/s49/s50/s52/s32/s75\n/s72\n/s97/s99/s61/s32/s51/s46/s56/s32/s79/s101/s32\n/s102/s32/s61/s32/s52/s57/s57/s32/s72/s122/s40/s100/s41 /s40/s99/s41\n/s32\n/s48/s72/s47/s47/s97/s98\n/s32\n/s48/s72/s47/s47/s99\n/s32/s32/s77 /s32/s40\n/s66/s47/s102/s46/s117/s46/s41\n/s48/s72/s32/s40/s84/s41/s84/s32/s61/s32/s50/s32/s75\nFIG. 3. (Color online) Temperature-dependent dc magnetic\nsusceptibility χ(T) in zero-field cooling (ZFC) and field cool-\ning (FC) modes taken at µ0H= 0.1 T for µ0H/bardblc(a) and\nµ0H/bardblab(inset), respectively. (b) 1 /χ(T) taken at µ0H=\n0.1 T along with Curie-Weiss fit from 150 to 300 K. (c) Field\ndependence of magnetization for FeCr 2Te4measured at T= 2\nK. (d) Ac susceptibility real part m′(T) measured with oscil-\nlating ac field of 3.8 Oe and frequency of 499 Hz. The chosen\nexperimental parameters (field and frequency) allow for wel l\ndefined moment and adequate frequency resolution.\nrange FM single crystals, such as U 2RhSi3[26]. The\nmagnetization loops of FeCr 2Te4for both field directions\natT=2Kconfirmsalargemagneticanisotropyandeasy\ncaxis [Fig. 3(c)]. The sudden jumps around µ0H≈ ±1\nTalongthe caxiscanbeascribedtothemagneticdomain\ncreeping behavior, i.e., the magnetic domain walls jump\nfrom one pinning site to another. Then we estimated the\nRhodes-Wohlfarth ratio (RWR) for FeCr 2Te4, which is\ndefined as Pc/PswithPcobtained from the effective mo-\nmentPc(Pc+2) =P2\neffandPsis the saturation moment\nobtained in the ordered state [27–29]. RWR is 1 for a lo-\ncalized system and is larger in an itinerant system. Here\nwe obtain RWR ≈1.69 for FeCr 2Te4, indicating a weak\nitinerant character. To obtain the accurate Curie tem-\nperature Tc, out-of-plane ac susceptibility was measured\nat oscillating ac field of 3.8 Oe and frequency of 499 Hz.\nThe single sharp peak in the real part m′(T) [Fig. 3(d)]\ngives the Tc= 124 K.\nIn the following we discuss the nature of the PM-FIM\ntransition for FeCr 2Te4. The magnetization isotherms\nalong easy caxis were measured at various temperatures\nin the vicinity of Tc[Fig. 4(a)]. We first considered the/s48 /s49 /s50 /s51 /s52 /s53/s48/s53/s49/s48/s49/s53/s50/s48/s50/s53/s51/s48\n/s48/s46/s48 /s48/s46/s49 /s48/s46/s50/s48/s50/s52/s54/s56\n/s49/s49/s48 /s49/s50/s48 /s49/s51/s48 /s49/s52/s48 /s49/s53/s48 /s49/s54/s48/s48/s46/s48/s48/s46/s52/s48/s46/s56/s49/s46/s50/s49/s46/s54/s50/s46/s48\n/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51/s48/s49/s50/s51/s52/s77/s32/s40/s101/s109/s117/s47/s103/s41\n/s109\n/s48/s72/s32/s40/s84/s41/s40/s97/s41\n/s84/s32/s61/s32/s49/s48/s48/s32/s75\n/s84/s32/s61/s32/s49/s53/s48/s32/s75\n/s68 /s84/s32/s61/s32/s50/s32/s75/s109\n/s48/s72/s32/s47/s47/s32/s99\n/s77/s50\n/s32/s40/s49/s48/s32/s101/s109/s117/s47/s103/s41/s50\n/s109\n/s48/s72/s47/s77/s32/s40/s84/s45/s103/s47/s101/s109/s117/s41/s40/s98/s41\n/s49/s53/s48/s32/s75/s49/s48/s48/s32/s75/s45/s68 /s83\n/s77/s32/s40/s74/s47/s107/s103/s45/s75/s41\n/s84/s32/s40/s75/s41/s109\n/s48/s72/s32/s40/s84/s41/s40/s100/s41\n/s53/s46/s48\n/s52/s46/s48\n/s51/s46/s53\n/s49/s46/s50/s49/s46/s54/s50/s46/s48/s50/s46/s53/s51/s46/s48/s52/s46/s53\n/s48/s46/s52/s32/s84/s48/s46/s56/s77/s49/s47/s98\n/s32/s40/s49/s48/s52\n/s32/s40/s101/s109/s117/s47/s103/s41/s49/s47 /s98\n/s41\n/s40/s109\n/s48/s72/s47/s77/s41/s49/s47 /s103\n/s32/s40/s84/s45/s103/s47/s101/s109/s117/s41/s49/s47 /s103/s40/s99/s41\n/s49/s48/s48/s32/s75\n/s49/s53/s48/s32/s75/s98 /s32/s61/s32/s48/s46/s51/s50\n/s103 /s32/s61/s32/s49/s46/s50/s49\nFIG. 4. (Color online) (a) Typical initial isothermal magne -\ntization curves from 100 K to 150 K with a temperature step\nof 2 K for FeCr 2Te4single crystal. (b) Arrott plot ( β= 0.5,\nγ= 1) and (c) the modified Arrott plot with the optimum ex-\nponents β= 0.32 andγ= 1.21. (d) Temperature-dependent\nmagnetic entropy change −∆SM(T) at various fields change.\nwell-known Arrott plot [30]. From the Landau theory,\nthe Arrott plot of M2vsH/Mshould appear as paral-\nlel straight lines above and below Tc, and the line passes\nthrough the origin at Tc. It is clear that the mean field\ncritical exponent does not work for FeCr 2Te4, as illus-\ntrated by the set of curved lines shown in Fig. 4(b).\nFor a second-order phase transition, the spontaneous\nmagnetization MsbelowTc, the inverse initial suscepti-\nbilityχ−1\n0aboveTc, and the field-dependent magnetiza-\ntionM(H) atTcare [31–33]:\nMs(T) =M0(−ε)β,ε <0,T < T c, (1)\nχ−1\n0(T) = (h0/m0)εγ,ε >0,T > T c,(2)\nM=DH1/δ,T=Tc, (3)\nwhereε= (T−Tc)/Tcis the reduced temperature, and\nM0,h0/m0andDare the critical amplitudes. In a\nmore general case, the modified Arrott plot ( H/M)1/γ=\naε+bM1/βwith self-consistent method was considered\n[34, 35]. Figure 4(c) presents the final modified Arrott\nplot ofM1/βvs (H/M)1/γwithβ= 0.32 andγ= 1.21,\nshowing a set of quasi-parallel lines at high field region.\nThen we extracted χ−1\n0(T) andMs(T) as the intercepts5\n/s50 /s51 /s52 /s53/s49/s54/s50/s52\n/s48 /s50 /s52 /s54 /s56 /s49/s48 /s49/s50 /s49/s52/s48/s50/s48/s52/s48/s54/s48/s56/s48\n/s45/s56 /s45/s52 /s48 /s52 /s56/s48/s49/s50/s51/s52/s49/s48/s56 /s49/s49/s52 /s49/s50/s48 /s49/s50/s54 /s49/s51/s50 /s49/s51/s56/s48/s52/s56/s49/s50/s49/s54/s50/s48\n/s84/s32/s40/s75/s41/s77\n/s115/s32/s40/s101/s109/s117/s47/s103/s41/s84\n/s99/s61/s32/s49/s50/s50/s46/s57/s40/s51/s41/s32/s75\n/s32/s61/s32/s48/s46/s51/s51/s40/s50/s41\n/s84\n/s99/s61/s32/s49/s50/s50/s46/s55/s40/s49/s41/s32/s75\n/s32/s61/s32/s49/s46/s50/s48/s40/s49/s41/s40/s97/s41\n/s48/s50/s52/s54/s56/s49/s48\n/s40/s49/s48/s45/s50\n/s32/s84/s45/s103/s47/s101/s109/s117/s41\n/s49/s48/s56 /s49/s49/s52 /s49/s50/s48 /s49/s50/s54 /s49/s51/s50 /s49/s51/s56/s45/s53/s48/s45/s52/s48/s45/s51/s48/s45/s50/s48/s45/s49/s48/s48\n/s84/s32/s40/s75/s41/s77\n/s115/s40/s100/s77\n/s115/s47/s100/s84/s41/s45/s49\n/s32/s40/s75/s41/s84\n/s99/s61/s32/s49/s50/s50/s46/s52/s40/s53/s41/s32/s75\n/s32/s61/s32/s48/s46/s51/s48/s40/s49/s41\n/s84\n/s99/s61/s32/s49/s50/s50/s46/s56/s40/s49/s41/s32/s75\n/s32/s61/s32/s49/s46/s50/s50/s40/s49/s41/s40/s98/s41\n/s48/s51/s54/s57/s49/s50\n/s40 /s100/s84/s41/s45/s49\n/s32\n/s40/s75/s41/s32/s61/s32/s52/s46/s56/s51/s40/s54/s41/s77/s32/s40 /s101/s109/s117/s32/s103/s45/s49\n/s41\n/s48/s72/s32/s40/s84/s41/s84\n/s99/s32/s61/s32/s49/s50/s51/s32/s75\n/s77 /s40/s101/s109/s117/s32/s103/s45/s49\n/s41\n/s48/s72 /s40 /s41/s84/s32/s60/s32/s84\n/s99\n/s84/s32/s62/s32/s84\n/s99/s40/s99/s41/s77/s40\n/s48/s72/s41/s45/s49/s47\n/s49/s48/s45/s51\n/s40\n/s48/s72/s41/s45/s84\n/s99/s32/s61/s32/s49/s50/s51/s32/s75/s40/s100/s41\nFIG. 5. (Color online) (a) Temperature dependence of the\nspontaneous magnetization Ms(left) and the inverse initial\nsusceptibility χ−1\n0(right)withsolid fittingcurves. Insetshows\nlogMvs log(µ0H) collected at Tc= 123 K with linear fitting\ncurve. (b) Kouvel-Fisher plots of Ms(dMs/dT)−1(left axis)\nandχ−1\n0(dχ−1\n0/dT)−1(right axis) with solid fitting curves.\n(c) Scaled magnetization mvs scaled field hbelow and above\nTcfor FeCr 2Te4. (d) The rescaling of the M(µ0H) curves by\nM(µ0H)−1/δvsε(µ0H)−1/(βδ).\non theH/Maxis and the positive M2axis, respectively.\nThe magnetic entropy change can be estimated using the\nMaxwell’s relation [36]:\n∆SM(T,H) =/integraldisplayH\n0/bracketleftbigg∂M(T,H)\n∂T/bracketrightbigg\nHdH. (4)\nFigure 4(d) presents the calculated −∆SMas a function\noftemperature. The −∆SMshowsabroadpeakcentered\nnearTcand the peak value monotonically increases with\nincreasing field. The maximum value of −∆SMreaches\n1.92 J kg−1K−1with a field change of 5 T. There is a\nslight shift of −∆SMpeak towards higher temperature\nwith increasing field, which also excludes the mean field\nmodel [37].\nFigure 5(a) presents the extracted Ms(T) andχ−1\n0(T)\nas a function of temperature. According to Eqs. (1) and\n(2), thecriticalexponents β= 0.33(2)with Tc= 122.9(3)\nK, andγ= 1.20(1) with Tc= 122.7(1) K, are obtained.\nThe exponent βdescribes the rapid increase of the or-\nder parameter below Tc. The exponent γdescribes how\nmagnetic susceptibility diverges at Tc. Here the obtained\nresult describes critical behavior of the net spontaneous\nmagnetization that arises in ferrimagnet. In the Kouvel-\nFisher (KF) relation [38]:\nMs(T)[dMs(T)/dT]−1= (T−Tc)/β, (5)χ−1\n0(T)[dχ−1\n0(T)/dT]−1= (T−Tc)/γ. (6)\nLinear fittings to the plots of Ms(T)[dMs(T)/dT]−1and\nχ−1\n0(T)[dχ−1\n0(T)/dT]−1in Fig. 5(b) yield β= 0.30(1)\nwithTc= 122.4(5) K, and γ= 1.22(1) with Tc=\n122.8(1) K. The third exponent δcan be calculated from\nthe Widom scaling relation δ= 1+γ/β[39]. From βand\nγobtainedwith the modified Arrottplotand the Kouvel-\nFisher plot, δ= 4.6(2) and 5.1(1) are obtained, respec-\ntively, which are close to the direct fit of δ= 4.83(6) tak-\ning into account that M=DH1/δatTc= 123K [inset in\nFig. 5(a)]. The obtained critical exponents of FeCr 2Te4\nare very close to the theoretically predicted values of 3D\nIsing model ( β= 0.325, γ= 1.24, and δ= 4.82) (Table\nII).\nScaling analysis can be used to estimate the reliability\nof the obtained critical exponents and Tc. The magnetic\nequation of state in the critical region is expressed as\nM(H,ε) =εβf±(H/εβ+γ), (7)\nwheref+forT > T candf−forT < T c, respectively,\nare the regular functions. Eq. (7) can be further written\nin terms of scaled magnetization m≡ε−βM(H,ε) and\nscaled field h≡ε−(β+γ)Hasm=f±(h). This suggests\nthat for true scaling relations and the right choice of β,\nγ, andδ, scaled mandhwill fall on universal curves\naboveTcand below Tc, respectively. As shown in Fig.\n5(c), allthe datacollapseon twoseparatebranchesbelow\nand above Tc, respectively. The scaling equation of state\ntakes another form,\nH\nMδ=k/parenleftBigε\nH1/β/parenrightBig\n, (8)\nwherek(x) is the scaling function. From the above equa-\ntion, all the data should also fall into a single curve.\nThis is indeed seen [Fig. 5(d)]; the M(µ0H)−1/δvs\nε(µ0H)−1/(βδ)experimental data collapse into a single\ncurve and the Tclocates at the zero point of the horizon-\ntal axis. The well-rescaled curves confirm the reliability\nof the obtained critical exponents and Tc.\nFurthermore, it is important to discuss the nature as\nwell as the range of magnetic interaction in FeCr 2Te4. In\na homogeneous magnet the universality class of the mag-\nnetic phase transition depends on the exchange distance\nJ(r). In renormalization group theory analysis the inter-\naction decays with distance rasJ(r)≈r−(3+σ), where\nσis a positive constant [41]. The susceptibility exponent\nγis:\nγ= 1+4\nd/parenleftbiggn+2\nn+8/parenrightbigg\n∆σ+8(n+2)(n−4)\nd2(n+8)2\n×/bracketleftBigg\n1+2G(d\n2)(7n+20)\n(n−4)(n+8)/bracketrightBigg\n∆σ2,(9)\nwhere ∆σ= (σ−d\n2) andG(d\n2) = 3−1\n4(d\n2)2,nis the spin\ndimensionality [42]. When σ >2, the Heisenberg model6\nTABLE II. Comparison of critical exponents of FeCr 2Te4with different theoretical models.\nReference Technique Tc− Tc+ β γ δ\nFeCr2Te4 This work Modified Arrott plot 122.9(3) 122.7(1) 0.33(2) 1.2 0(1) 4.6(2)\nThis work Kouvel-Fisher plot 122.4(5) 122.8(1) 0.30(1) 1.2 2(1) 5.1(1)\nThis work Critical isotherm 4.83(6)\n3D Heisenberg 28 Theory 0.365 1.386 4.8\n3D XY 28 Theory 0.345 1.316 4.81\n3D Ising 28 Theory 0.325 1.24 4.82\nTricritical mean field 36 Theory 0.25 1.0 5.0\nTABLE III. The first-principles total energy (in meV) per for -\nmula unit of different magnetic patterns as shown in Fig. 6(a)\nusing the 300 K and 105 K experimental structures. In the\nE-AF-2 pattern, a Cr atom is antiferromagnetically and fer-\nromagnetically aligned with the closest and farthermost of its\nsix nearest Cr neighbors, respectively; the opposite holds in\nthe E-AF-1 pattern. FM means the ferromagnetic configura-\ntion.\nStructure Ferri FM C-AF E-AF-1 E-AF-2\n300 K 8 90 197 31 82\n105 K 0 56 171 72 32\nis valid forthe 3Disotropicmagnet, where J(r) decreases\nfaster than r−5. Whenσ≤3/2, the mean-field model is\nsatisfied, expectingthat J(r) decreasesslowerthan r−4.5.\nFor the 3D-Ising model with d= 3 and n= 1,σ= 1.88\nis obtained, leading to spin interactions J(r) decaying as\nJ(r)≈r−4.88. This calculation suggests that the spin\ninteraction in FeCr 2Te4is close to the 3D Ising localized-\ntype coupled with a long-range ( σ= 1.88) interaction,\nin line with its weak itinerant character. Meanwhile, the\ncorrelation length ( ξ) correlates with the critical expo-\nnentν(ν=γ/σ), where ξ=ξ0[(T−Tc)/Tc]−ν. It gives\nthatν= 0.64(1) and α= 0.08 (α= 2−νd).\nTo get further insight into the magnetism, we per-\nformed first-principles calculations using density func-\ntion theory. We applied the WIEN2K implementation\n[43] of the full potential linearized augmented plane-wave\nmethod in generalized-gradient approximation using the\nPBEsolfunctional [44]. The basis size wasdetermined by\nRmtKmax= 7 and the Brillouin zone was sampled with\n115 irreducible kpoints to achieve energy convergence of\n1 meV. As shown in Table III, we found that for the ex-\nperimental structure refined at 105 K, the ferrimagnetic\nstate where the Cr and Fe atoms have opposite spin ori-\nentationsis56meVperformulaunitlowerintotalenergy\nthan the ferromagnetic phase and 32 meV lowered than\nthe most stable antiferromagnetic structure (i.e., E-AF-\n2) as observed in FeCr 2Se4[45]. A similar trend of the\nresults holds for the the experimental structure refined\nat 300 K (Table III). A weak easy c-axis anisotropy was\nobtained by inclusion of spin-orbit coupling in the calcu-\nFIG. 6. (Color online) (a) The antiferromagnetic structure s\nused in the first-principles total energy calculations. Ato m-\nresolved density of states (b) in the nonmagnetic state and\n(c) in the ferrimagnetic state where the Cr and Fe atoms\nhave opposite spin orientations, which are shown in the rati o\nof two Cr ions to one Fe ion.\nlations, namely the total energy per formula unit is lower\nby less than 1 meV for the magnetizationalong the caxis\nthan along the aorbaxis. Thus, other sources of mag-\nnetic anisotropy such as dipole-dipole interaction are im-\nportant. The calculated atom-resolved density of states\n(DOS) is shown in Figs. 6(b) and 6(c). For the nonmag-\nnetic case, the Cr-derived DOS has a sharp peak at the\nFermi level [Figs. 6(b)], suggesting a strong Stoner insta-\nbility that yields an itinerant ferromagnetism in the Cr\nlayers. This yields the splitting of about 2.8 eV between\nthe spin-majority and spin-minority bands of Cr charac-7\nter with the dramatic reduction of the Cr-derived DOS\nat the Fermi level [Fig. 6(c)], indicative of the localized\nspin picturefor the Cratoms. Whereas, the nonmagnetic\nFe-derived DOS peaks at about 0 .6 eV below the Fermi\nlevel [Fig. 6(b)] and experiences little reduction at the\nFermi level upon entering the magnetic phase [Fig. 6(c)].\nWe infer that the magnetism of the Fe ions is established\nvia antiferromagnetic superexchange with the neighbor-\ning two Cr ions. The Cr magnetic moment within the\natomic Muffin tins is about 2.88 µB, which is close to\nthe nominal Cr3+S= 3/2 state. With the octahedral\ncoordination, the splitting of the five 3 dorbitals between\nthe high-lying egand low-lying t2gorbitals is substantial.\nTheS= 3/2 state of the Cr3+ion (i.e., 3 d3ort3\n2ge0\ngelec-\ntronconfiguration)meansthattheCr t2gorbitalsarehalf\nfilled, rendering a vanishing orbital angular moment and\na negligible spin-orbit coupling effect. The Fe magnetic\nmoment is about 2.80 µB, significantly deviated from the\nnominal high-spin Fe2+S= 2 state, which indicates its\ndual characters with both localized spins and itinerant\nelectrons. This reveals an interesting interplay of Cr and\nFeelectronicstates,whichallowsthespin-majoritybands\nof the system at the Fermi level to be of Fe character\nrather than Cr character [Fig. 6(c)]. We thus picture\nthe FIM in FeCr 2Te4as itinerant ferromagnetism among\nthe antiferromagnetically coupled Cr-Fe-Cr trimers. The\ntrimers centered at the Fe sites form a body-centered\northorhombic lattice of magnetic dipoles with effective\nmoment of 2 µBS= 4µB[Fig. 3(c)]. In addition to the\neffective Heisenberg exchange interaction, the ith trimer\nis coupled with its ten neighboring trimers [Fig. 6(a)]\nvia dipole-dipole interaction ∝ −(Si·rij)(Sj·rij)/|rij|5,\nwhich tends to align the magnetic moments along the\nbond direction rij=ri−rjwherejdenotes one of the\nneighboring trimers and riis the spatial vector of the ith\ntrimer. Since the body-centered orthorhombic structure\nof the trimers is substantially elongated along the caxis,\neasyc-axis magnetic anisotropy has the overall minimum\ndeviation from the neighboring bond directions. We thus\npredict that the 3D Ising-like ferrimagnetism is sensitive\nto changes in the lattice structure, especially the tilting\nof the Cr-Fe-Cr trimers, which will be verified by future\npressure experiments and computer simulations.\nCONCLUSIONS\nIn summary, we systematically investigated structural\nand magnetic properties of stoichiometric FeCr 2Te4that\ncrystallizes in the I2/mspace group. The second-order\nPM-FIM transition is observed at Tc∼123 K. The\ncritical exponents β,γ, andδestimated from various\ntechniques match reasonably well and follow the scaling\nequation. The analysis of critical behavior suggests that\nFeCr2Te4isa3D-Isingsystemdisplayingalong-rangeex-\nchange interaction with the exchange distance decayingasJ(r)≈r−4.88. Combined experimental and theoret-\nical analysis attributes the ferrimagnetism in FeCr 2Te4\nto itinerant ferromagnetism among the antiferromagneti-\ncallycoupledCr-Fe-Crtrimers. Follow-upstudiesoflocal\natomic structure and magnetism using x-rayand neutron\nscatteringaswellashigh-pressuremethodswillbeofpar-\nticular interest for more comprehensive understanding of\nthis system.\nACKNOWLEDGEMENTS\nWork at BNL is supported by the Office of Basic En-\nergy Sciences, Materials Sciences and Engineering Divi-\nsion, U.S. Department of Energy (DOE) under Contract\nNo. DE-SC0012704. This research used the 28-ID-1 and\n8-ID beamlines of the NSLS II, a U.S. DOE Office of Sci-\nence User Facility operated for the DOE Office of Science\nby BNL under Contract No. DE-SC0012704. This re-\nsearch used resources of the Center for Functional Nano-\nmaterials (CFN), which is a U.S. DOE Office of Science\nFacility, at BNL under Contract No. DE-SC0012704.\n[1] T. Kanomata, H. Ido, and T. Kaneko, J. Phys. Soc. Jpn.\n29, 332 (1970).\n[2] H. W. Lehmann, and F. P. Emmenegger, Solid State\nCommun. 7, 965 (1969).\n[3] N. Menyuk, K. Dwight, and R. J. Arnott, J. Appl. Phys.\n37, 1387 (1966).\n[4] Z. R. Yang, S. Tan, Z. W. Chen, and Y. H. Zhang, Phys.\nRev. B62, 13872 (2000).\n[5] V. Tsurkan, M. M¨ ucksch, V. Fritsch, J. Hemberger, M.\nKlemm, S. Klimm, S. K¨ orner, H. A. Krug von Nidda, D.\nSamusi, E. W.Scheidt, A.Loidl, S.Horn, andR.Tidecks,\nPhys. Rev. B 68, 134434 (2003).\n[6] K. Ohgushi, Y. Okimoto, T. Ogasawara, S. Miyasaka,\nand Y. Tokura, J. Phys. Soc. Jpn. 77, 034713 (2019).\n[7] J.Hemberger, T.Rudolf, H.A.KrugvonNidda, F.Mayr,\nA. Pimenov, V. Tsurkan, and A. Loidl, Phys. Rev. Lett.\n97, 087204 (2006).\n[8] J. Bertinshaw, C. Ulrich, A. Gunter, F. Schrettle, M.\nWohlauer, S. Krohns, M. Reehuis, A.J. Studer, M.\nAvdeev, D.V. Quach, J.R. Groza, V. Tsurkan, A. Loidl,\nand J. Deisenhofer, Sci. Rep. 4, 6079 (2014).\n[9] V. Tsurkan, O. Zaharko, F. Schrettle, C. Kant, J. Deisen-\nhofer, H. A. Krug von Nidda, V. Felea, P. Lemmens, J.\nR. Groza, D. V. Quach, F. Gozzo, and A. Loidl, Phys.\nRev B81, 184426 (2010).\n[10] V. Tsurkan, I. Fita, M. Baran, R. Puzniak, D.Samusi,\nR.Szymczak, H. Szymczak, S. Lkimm, M.Kliemm, S.\nHonn, and R. Tidecks, J. Appl. Phys. 90, 875 (2001).\n[11] A. P.Ramirez, R.J. Cava, andJ. Krajewski, Nature 386,\n156 (1997).\n[12] B. I. Min, S. S. Abik, H. C. Choi, S. K. Kwon, and J. S.\nKang, New Jour. Phys. 10, 055014 (2008).\n[13] G. J. Snyder, T. Caillat, and J. P. Fleurial, Phys. Rev.\nB62, 10185 (2000).8\n[14] H. N. Ok, and C. S. Lee, Phys. Rev. B 33, 581 (1986).\n[15] J. S. Kang, G. Kim, H. J. Lee, H. S. Kim, D. H. Kim, S.\nW. Han, S. J. Kim, C. S. Kim, H. Lee, J. Y. Kim, and\nB. I. Min, J. Appl. Phys. 103, 07D717 (2008).\n[16] A. B. Demeaux, G. Villers, and P. Gibart, J. Solid State\nChem.15, 178 (1975).\n[17] L. M. Valiev, I. G. Kerimov, S. Kh. Dabaev, and Z. M.\nNamazov, Inorg. Mater. 11, 213 (1975).\n[18] C. S. Yadav, S. K. Pandey, and P. L. Paulose, arXiv:\n1904.06661.\n[19] E.S. Bozin, W.G. Yin, R.J. Koch, M. Abeykoon, Y.S.\nHor, H. Zheng, H.C. Lei, C. Petrovic, J.F. Mitchell and\nS.J.L. Billinge, Nature Comms. 10, 3638 (2019).\n[20] J. Kieffer and J. P. Wright, Powder Diffraction 28, 339\n(2013).\n[21] B. H. Toby and R. B. Von Dreele, J. Appl. Crystallogr.\n46, 544 (2013).\n[22] A. Ignatov, C. L. Zhang, M. Vannucci, M. Croft,\nT. A. Tyson, D. Kwok, Z. Qin, and S. W. Cheong,\narXiv:0808.2134v2.\n[23] H.Ofuchi, N.Ozaki, N.Nishizawa, H. Kinjyo, S. Kuroda,\nand K. Takita, AIP Conference Proceeding 882, 517\n(2017).\n[24] R. Prins and D. C.Koningsberger (eds.), X-ray Absorp-\ntion: Principles, Applications, Techniques of EXAFS,\nSEXAFS, XANES (Wiley, New York, 1988).\n[25] K. Adachi, K. Sato, and K. Kojima, Mem. Fac. Eng.\nNagoya Univ. Jpn. 22, 253 (1970).\n[26] M. Szlawska, M. Majewicz, and D. Kaczorowski, J. Al-\nloys Compd. 662, 208 (2016).\n[27] E. P. Wohlfarth, J. Magn. Magn. Mater., 7, 113 (1978).\n[28] T. Moriya, J. Magn. Magn. Mater., 14, 1 (1979).\n[29] Y. Takahashi, Spin Fluctuation Theory of Itinerant Elec-\ntron Magnetism (Springer Tracts in Modern Physics, vol.253, Springer-Verlag, Berlin, Heidelberg, 2013).\n[30] A. Arrott, Phys. Rev. B 108, 1394 (1957).\n[31] H. E. Stanley, Introduction to Phase Transitions and\nCritical Phenomena (Oxford U. P., London and New\nYork, 1971).\n[32] M. E. Fisher, Rep. Prog. Phys. 30, 615 (1967).\n[33] J. Lin, P. Tong, D. Cui, C. Yang, J. Yang, S. Lin, B.\nWang, W. Tong, L. Zhang, Y. Zou, and Y. Sun, Sci.\nRep.5, 7933 (2015).\n[34] W. Kellner, M. F¨ ahnle, H. Kronm¨ uller, and S. N. Kaul,\nPhys. Status Solidi B 144, 387 (1987).\n[35] A. K. Pramanik, and A. Banerjee, Phys. Rev. B 79,\n214426 (2009).\n[36] J. Amaral, M. Reis, V. Amaral, T. Mendonc, J. Araujo,\nM. Sa, P. Tavares, J. Vieira, J. Magn. Magn. Mater. 290,\n686 (2005).\n[37] V. Franco, A. Conde, M. D. Kuzmin, and J. M. Romero-\nEnrique, J. Appl. Phys. 105, 07A917 (2009).\n[38] J. S. Kouvel, and M. E. Fisher, Phys. Rev. 136, A1626\n(1964).\n[39] B. Widom, J. Chem. Phys. 41, 1633 (1964).\n[40] S. Kaul, J. Magn. Magn. Mater. 53, 5 (1985).\n[41] M. E. Fisher, S. K. Ma, and B. G. Nickel, Phys. Rev.\nLett.29917 (1972).\n[42] S. F. Fischer, S. N. Kaul, and H. Kronmuller, Phys. Rev.\nB65, 064443 (2002).\n[43] K. Schwarz, P. Blaha, and G. K. H. Madsen, Comput.\nPhys. Commun. 147, 71 (2002).\n[44] J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov,\nG. E. Scuseria, L. A. Constantin, X. Zhou, and K. Burke,\nPhys. Rev. Lett. 100, 136406 (2008).\n[45] S. R. Hong and H. N. Ok, Phys. Rev. B 76, 4176 (1975)." }, { "title": "1712.05622v1.Effect_of_the_Canting_of_Local_Anisotropy_Axes_on_Ground_State_Properties_of_a_Ferrimagnetic_Chain_with_Regularly_Alternating_Ising_and_Heisenberg_Spins.pdf", "content": "arXiv:1712.05622v1 [cond-mat.str-el] 15 Dec 2017Vol.XXX (201X) CSMAG‘16 No.X\nEffect of the Canting of Local Anisotropy Axes on Ground-Stat e Properties of a\nFerrimagnetic Chain with Regularly Alternating Ising and H eisenberg Spins\nJ. Torrico,1,∗M.L. Lyra,1O. Rojas,2S.M. de Souza,2and J. Streˇ cka3\n1Instituto de F´ ısica, Universidade Federal de Alagoas, 570 72-970 Maceio, AL, Brazil\n2Departamento de F´ ısica, Universidade Federal de Lavras, 3 7200-000, Lavras-MG\n3Institute of Physics, Faculty of Science, P. J. ˇSaf´ arik University, Park Angelinum 9, 040 01 Koˇ sice, Slov akia\nThe effect of the canting of local anisotropy axes on the groun d-state phase diagram and mag-\nnetization of a ferrimagnetic chain with regularly alterna ting Ising and Heisenberg spins is exactly\nexamined in an arbitrarily oriented magnetic field. It is sho wn that individual contributions of Ising\nand Heisenberg spins to the total magnetization basically d epend on the spatial orientation of the\nmagnetic field and the canting angle between two different loc al anisotropy axes of the Ising spins.\nPACS numbers: 75.10.Pq ; 75.10.Kt ; 75.30.Kz ; 75.40.Cx ; 75. 60.Ej\nIntroduction\nIn spite of a certain over-simplification, a few\nexactly solved Ising-Heisenberg models capture es-\nsential magnetic features of some real polymeric\ncoordination compounds as for instance Cu(3-\nClpy)2(N3)2[1], [(CuL) 2Dy][Mo(CN) 8] [2, 3] and\n[Fe(H2O)(L)][Nb(CN) 8][Fe(L)] [4]. The rigorous solu-\ntions for the Ising-Heisenberg models thus afford an\nexcellent playground for experimental testing of a lot\nof intriguing magnetic properties such as quantized\nmagnetization plateaus, anomalous thermodynamics,\nenhanced magnetocaloric effect, etc. [1–4]\nRecently, it has been verified that the bimetallic coor-\ndination polymer Dy(NO 3)(DMSO) 2Cu(opba)(DMSO) 2\n(to be further abbreviated as DyCu) can be satisfactorily\ndescribed by the spin-1/2 Ising-Heisenberg chain with\nregularly alternating Ising and Heisenberg spins, which\ncapture the magnetic behavior of Dy3+and Cu2+mag-\nnetic ions, respectively [5]. However, a closer inspection\nof available structural data reveals two crystallographi-\ncally inequivalent orientantions of coordination polyhe-\ndra of Dy3+magnetic ions, which regularly alternate\nalong the DyCu chain [6]. Motivated by this fact, we\nwill investigate in the present work the effect of the cant-\ning between two different local anisotropy axes on the\nground-state properties of the spin-1/2 Ising-Heisenberg\nchain with regularly alternating Ising and Heisenberg\nspins in an arbitrarily oriented magnetic field.\nModel and its Hamiltonian\nLet us introduce the spin-1/2 Ising-Heisenberg chain\nschematically illustrated in Fig. 1, in which the Ising\n∗corresponding author; e-mail: jordanatorrico@gmail.comspins with two different local anisotropy axes z1andz2\nregularly alternate with the Heisenberg spins. The local\nanisotropy axis z1(z2) of the Ising spins σ= 1/2 on odd\n(even) lattice positions is canted by the angle α(-α) from\nthe global frame z-axis. Hence, it follows that the angle\n2αdetermines the overall canting between two coplanar\nlocal anisotropy axes z1andz2. The Heisenberg spins\nS= 1/2 are coupled to their nearest-neighbor Ising spins\nthrough the antiferromagnetic coupling J <0 projected\nintothe respectiveanisotropyaxis. Furthermore, we take\ninto account the effect of the external magnetic field B,\nwhose spatial orientation is given by the angle θdeter-\nmining its tilting from the global frame z-axis. Under\nthese circumstances, the spin-1/2 Ising-Heisenberg chain\ncan be defined through the following Hamiltonian\nH=−JN/2/summationdisplay\ni=1(Sz1\n2i−1σz1\n2i−1+Sz2\n2i−1σz2\n2i+Sz2\n2iσz2\n2i+Sz1\n2iσz1\n2i+1)\n−hz1N/2/summationdisplay\ni=1σz1\n2i−1−hz2N/2/summationdisplay\ni=1σz2\n2i\n−hzN/summationdisplay\ni=1Sz\ni−hxN/summationdisplay\ni=1Sx\ni, (1)\nwherehz1=gz1\n1µBBcos(α−θ) andhz2=gz2\n1µBBcos(α+\nθ) determine projections of the external magnetic field B\ntowards the anisotropy axes of the Ising spins on odd\nand even lattice positions, respectively, gz1\n1andgz2\n1are\nthe respective Land´ e g-factors of the Ising spins and µB\nis the Bohr magneton. Similarly, hz=gz\n2µBBcosθand\nhx=gx\n2µBBsinθdetermine two orthogonal projections\nof the external magnetic field for the Heisenberg spins,\nwhereasgz\n2andgx\n2are the respective spatial components\nof the Land´ e g-factors of the Heisenberg spins.\nThetotalHamiltonianofthespin-1/2Ising-Heisenberg\nchaincanberewrittenasthesumofthecellHamiltonians\nH=N/2/summationdisplay\ni=1(H2i−1+H2i), (2)Ferrimagnetic Chain of Alternating Ising and Heisenberg Sp ins 2\nz z z z z\nxyz α α αz1 z1 z1 z2z2/vectorBθ\nσ2i−1σ2iS2i S2i−1\nSi= 1 /2 Heisenberg spins σi= 1 /2 Ising spins α′=−αα′α′\nFIG. 1: (Color online) Schematic representation of a spin\nchain with regularly alternating Ising and Heisenberg spin s.\nThe angle α(-α) determines the canting of the local\nanisotropy axis z1(z2) from the global frame z-axis for odd\n(even) Ising spins so that 2 αis the canting angle between two\ncoplanar anisotropy axes. The angle θdetermines the tilting\nof the magnetic field from the global frame z-axis.\neach of which involves all the interaction and field terms\nof exactly one Heisenberg spin\nH2i−1=−hz1\n2σz1\n2i−1−hz2\n2σz2\n2i−hz\n2i−1Sz\n2i−1−hx\n2i−1Sx\n2i−1,\nH2i=−hz2\n2σz2\n2i−hz1\n2σz1\n2i+1−hz\n2iSz\n2i−hx\n2iSx\n2i.(3)\nIn above, we have introduced the following notation for\nthe effective longitudinal and transverse fields acting on\nthe Heisenberg spins\nhz\n2i−1=Jcosα/parenleftbig\nσz1\n2i−1+σz2\n2i/parenrightbig\n+gz\n2µBBcosθ,\nhz\n2i=Jcosα/parenleftbig\nσz2\n2i+σz1\n2i+1/parenrightbig\n+gz\n2µBBcosθ,\nhx\n2i−1=Jsinα/parenleftbig\nσz1\n2i−1−σz2\n2i/parenrightbig\n+gx\n2µBBsinθ,\nhx\n2i=−Jsinα/parenleftbig\nσz2\n2i−σz1\n2i+1/parenrightbig\n+gx\n2µBBsinθ.(4)\nIt is noteworthy that the cell Hamiltonians (3) commute\nand hence, they can be diagonalized independently of\neach other by performing a local spin-rotation transfor-\nmation following the approach worked out previously [5].\nIn this way, one obtains the full spectrum of the eigenval-\nues, which can be subsequently utilized for the construc-\ntion of the ground-state phase diagram and magnetiza-\ntionprocess. Thefulldetailsofthiscalculationprocedure\nwill be published elsewheretogetherwith a morecompre-\nhensive analysis of the thermodynamic properties.\nResults and discussion\nLet us illustrate a few typical ground-state phase di-\nagrams and zero-temperature magnetization curves for\nthe most interesting particular case with the antiferro-\nmagnetic coupling J <0, equal Land´ e g-factors of the\nIsing spins gz1\n1=gz2\n1= 20 and equal components of the\nLand´ e g-factor of the Heisenberg spins gx\n2=gz\n2= 2,\nwhich nearly coincide with usual values of gyromagnetic\nratioforDy3+andCu2+magneticions, respectively. Un-\nder these circumstances, one finds four different ground\nstates: two ground states CIF 1and CIF 2with the cantedferromagnetic alignment of the Ising spins\n|CIF1/angbracketright=N/2/productdisplay\ni=1| ր/angbracketright2i−1|ψ/angbracketright2i−1| տ/angbracketright2i|ψ/angbracketright2i,(5)\n|CIF2/angbracketright=N/2/productdisplay\ni=1| ւ/angbracketright2i−1|ψ/angbracketright2i−1| ց/angbracketright2i|ψ/angbracketright2i,(6)\nand two ground states CIA 1and CIA 2with the canted\nantiferromagnetic alignment of the Ising spins\n|CIA1/angbracketright=N/2/productdisplay\ni=1| ր/angbracketright2i−1|ψ/angbracketright2i−1| ց/angbracketright2i|ψ/angbracketright2i,(7)\n|CIA2/angbracketright=N/2/productdisplay\ni=1| ւ/angbracketright2i−1|ψ/angbracketright2i−1| տ/angbracketright2i|ψ/angbracketright2i.(8)\nIt is noteworthythat the state vector | ր/angbracketright2i−1(| ւ/angbracketright2i−1)\ncorrespondstothespinstate σz1\n2i−1= 1/2(σz1\n2i−1=−1/2)\noftheodd-siteIsingspins, thestatevector | տ/angbracketright2i(| ց/angbracketright2i)\ncorresponds to the spin state σz2\n2i= 1/2 (σz2\n2i=−1/2)\nof the even-site Ising spins, while each Heisenberg spin\nunderlies a quantum superposition of both spin states\n|ψ/angbracketrighti=1/radicalbig\na2\ni+1(| ↓/angbracketrighti−ai| ↑/angbracketrighti), (9)\nwhich depends on the orientation of its two nearest-\nneighbor Ising spins via ai=hx\ni/[hz\ni−/radicalbig\n(hz\ni)2+(hx\ni)2].\nFIG. 2: (Color online) The ground-state phase diagram in\npolar coordinates for two different canting angles between t he\nlocal anisotropy axes: (a) 2 α=π/6; (b) 2α=π/4. The\nrelative size of the magnetic field µBB/|J|is represented by\nthe radius of the polar coordinates and the angle θdetermines\nits inclination with respect to the global frame z-axis.\nThe overall ground-state phase diagram in polar co-\nordinates is illustrated in Fig. 2 for two different canting\nangles 2αbetween both local anisotropyaxes. The phase\ndiagram has an obvious symmetry with respect to θ= 0\nandπ/2 axes, the former symmetry axis θ= 0 merely\ninterchanges CIA 1↔CIA2, while the latter symmetry\naxisθ=π/2 is responsible for CIF 1↔CIF2interchange.\nWithout loss of generality, our further discussion will be\ntherefore restricted just to the first quadrant θ∈[0,π/2].\nThe coexistence line between CIF 1and CIA 1phases3 Ferrimagnetic Chain of Alternating Ising and Heisenberg Sp ins\nis macroscopically degenerate with the residual entropy\nper Ising-Heisenberg pair S=kBln(2)/2, whereas CIF 1\nand CIF 2phases coexist together at θ=π/2 up to a\ntriple point (diamond symbol) with the residual entropy\nS=kBln[(√\n5 + 3)/2]/2. In addition, the Heisenberg\nspinsarecompletelyfreetoflip atmacroscopicallydegen-\nerate points (blue circles) with the residual entropy S=\nkBln(2) given by the coordinates B=|J|cos(α)/(2µB),\nθ= 0 andB=|J|sin(α)/(2µB),θ=π/2 forα>∼π/9.\nThese highly degenerate points correspond to a novel-\ntype spin frustration ’half ice, half fire’ [7], which origi-\nnates from the difference between Land´ e g-factors being\nresponsible for a fully frozen (ordered) character of the\nIsing spins and fully floppy (disordered) character of the\nHeisenberg spins.\nFinally, the individual contributions of the Ising and\nHeisenberg spins to the total magnetization are depicted\nin Fig. 3 as a function of the magnetic-field strength for\nseveral spatial orientations θof the applied field. As one\ncan see, any deviation of the magnetic field from its lon-\ngitudinal direction θ= 0 destroys the sharp stepwise de-\npendence in the longitudinalprojectionofthe magnetiza-\ntionmz\n2of the Heisenberg spins. In fact, the longitudinal\ncomponent mz\n2is gradually smeared out upon increasing\nofthe tiltingangle θ, whileits transversepart mx\n2risesup\nto its global maximum successively followed by a gradual\ndecline [cf. Fig. 3(a)-(b)]. The most notabledependences\nofthemagnetizationoftheHeisenbergspinscanbefound\nfor greater tilting angles θof the magnetic field, which\ncause an abrupt jump in both components mx\n2andmz\n2\nof the Heisenberg spins due to a sudden reorientation of\nIsing spins at the phase transition from the canted fer-\nromagnetic phase CIF 1to the canted antiferromagnetic\nphase CIA 1(see the curves for θ= 2π/5). The sharp\nstepwise dependence of the magnetization of the Heisen-\nberg spins is afterwards recovered for the special case of\nthe transverse field θ=π/2, for which it appears in the\ntransverse projection mx\n2while its longitudinal part mz\n2\nequals zero. The coexistence of the canted ferromagnetic\nphases CIF 1and CIF 2is manifested through a quasi-\nlinear dependence of mx\n2at low magnetic fields, where\nother contributions mz\n2=mz1\n1=mz2\n1= 0 vanish.\nConclusion\nIn the present work we have examined in detail the\nground-state phase diagram and zero-temperature mag-\nnetization process of the spin-1/2 Ising-Heisenberg chain\nwith two different local anisotropy axes in an arbitrar-\nily oriented magnetic field. It has been shown that the\nphase diagram involvesin total two canted ferromagnetic\nand two canted antiferromagnetic ground states. An-\nother interesting finding concerns with the existence of a\nfewmacroscopicallydegeneratepoints, atwhichaperfect\norder of the Ising spins accompanies a complete disorder/s48/s46/s48 /s48/s46/s52 /s48/s46/s56 /s49/s46/s50/s45/s48/s46/s53/s48/s45/s48/s46/s50/s53/s48/s46/s48/s48/s48/s46/s50/s53/s48/s46/s53/s48\n/s48/s46/s48 /s48/s46/s52 /s48/s46/s56 /s49/s46/s50/s45/s48/s46/s53/s48/s45/s48/s46/s50/s53/s48/s46/s48/s48/s48/s46/s50/s53/s48/s46/s53/s48\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s45/s48/s46/s53/s48/s45/s48/s46/s50/s53/s48/s46/s48/s48/s48/s46/s50/s53/s48/s46/s53/s48/s109/s120 /s50/s32/s32/s91/s103/s120 /s50\n/s66/s93\n/s32/s32\n/s32/s32 /s32\n/s32 /s32\n/s32\n/s40/s98/s41\n/s32/s32\n/s32/s109/s122 /s50/s32/s32/s91/s103/s122 /s50\n/s66/s93\n/s40/s99/s41\n/s32/s32/s109/s122\n/s49\n/s49/s32/s32/s32/s32/s91/s103/s122\n/s49\n/s49 /s66/s93\n/s66/s47/s124/s74/s124/s40/s97/s41\n/s40/s100/s41\n/s32/s32/s109/s122\n/s50\n/s49/s32/s32/s32/s32/s91/s103/s122\n/s50\n/s49 /s66/s93\n/s66/s47/s124/s74/s124\nFIG. 3: (Color online) Zero-temperature magnetizations ve r-\nsus the relative strength of the magnetic field for the cantin g\nangle 2α=π/4 and several spatial orientations θof the ap-\nplied field: (a)-(b) the transverse mx\n2and longitudinal mz\n2\nprojections of the Heisenberg spins; (c)-(d) the local proj ec-\ntionsmz1\n1andmz2\n1of the Ising spins towards their easy axes.\nof the Heisenberg spins within the so-called ’half ice, half\nfire’ frustrated ground state [7]. It has been also convinc-\ningly evidenced that the canting angle between two local\nanisotropy axes of the Ising spins and the spatial orien-\ntation of the applied magnetic field basically influences\nthe overall shape of the magnetization curves.\nAcknowledgments\nThis work was partially supported by FAPEAL\n(Alagoas State Research agency), CNPq, CAPES,\nFAPEMIG, VEGA 1/0043/16 and APVV-14-0073.\n[1] J. Streˇ cka, M. Jaˇ sˇ cur, M. Hagiwara, K. Minami, Y.\nNarumi, K. Kindo, Phys. Rev. B 72, 024459 (2005).\nDOI:10.1103/PhysRevB.72.024459.\n[2] W. Van den Heuvel, L.F. Chibotaru, Phys. Rev. B 82,\n174436 (2010). DOI:10.1103/PhysRevB.82.174436.\n[3] S. Bellucci, V. Ohanyan, O. Rojas, EPL105, 47012\n(2014). DOI: 10.1209/0295-5075/105/47012\n[4] S. Sahoo, J.P. Sutter, S. Ramasesha, J. Stat. Phys. 147,\n181 (2012). DOI: 10.1007/s10955-012-0460-7.\n[5] J. Streˇ cka, M. Hagiwara, Y. Han, T. Kida, Z. Honda,\nM. Ikeda, Condens. Matter Phys. 15, 43002 (2012). DOI:\n10.5488/CMP.15.43002.\n[6] G. Calvez, K. Bernot, O. Guillou et al., Inorg. Chim. Acta\n361, 3997 (2008). DOI: 10.1016/j.ica.2008.03.040.\n[7] W. Yin, Ch. Roth, A. Tsvelik, arxiv: 1510.00030v2." }, { "title": "1807.02445v1.Spin_torque_induced_magnetization_dynamics_in_ferrimagnets_based_on_Landau_Lifshitz_Bloch_Equation.pdf", "content": "Spin-torque-induced magnetization dynamics in ferrimagnets based on\nLandau-Lifshitz-Bloch Equation\nZhifeng Zhu,1,\u0003Xuanyao Fong,1and Gengchiau Liang1,y\n1Department of Electrical and Computer Engineering,\nNational University of Singapore, Singapore 117576\n(Dated: July 9, 2018)\nA theoretical model based on the Landau-Lifshitz-Bloch equation is developed to study the spin-\ntorque e\u000bect in ferrimagnets. Experimental \fndings, such as the temperature dependence, the peak\nin spin torque, and the angular-momentum compensation, can be well captured. In contrast to\nthe ferromagnet system, the switching trajectory in ferrimagnets is found to be precession free.\nThe two sublattices are not always collinear, which produces large exchange \feld a\u000becting the\nmagnetization dynamics. The study of material composition shows the existence of an oscillation\nregion at intermediate current density, induced by the nondeterministic switching. Compared to the\nLandau-Lifshitz-Gilbert model, our developed model based on the Landau-Lifshitz-Bloch equation\nenables the systematic study of spin-torque e\u000bect and the evaluation of ferrimagnet-based devices.\nI. INTRODUCTION\nFerrimagnets (FiMs) with antiferromagnetic exchange\ncoupled transition-metal (TM) and rare-earth (RE) al-\nloys have attracted considerable attention due to the rich\nphysics1{7and their promise in device applications8{10.\nThe FiMs are expected to have fast spin dynamics like\nantiferromagnets (AFMs), but their magnetic states can\nbe electrically sensed using the tunnel magnetoresistance\n(TMR) e\u000bect due to the \fnite net magnetization ( mnet),\nwhich can be tuned by temperature ( T) or material com-\nposition (X). In addition, the FiMs have large bulk\nperpendicular anisotropy, which o\u000bers an alternative to\nthe ferromagnets (FMs) and enables the scaling down of\nMRAM down to 20 nm8. Furthermore, di\u000berent g factors\nbetween sublattices induce an angular-momentum com-\npensation point, which enables fast domain-wall motion9.\nThe FiMs can be manipulated by magnetic \feld or\nlaser heating2,3,11{15, but an electrical method, such as\nthe spin-transfer torque (STT)1or the spin-orbit torque\n(SOT)4,16, is preferred for electrical characterizations\nand applications. Therefore, it is important to study the\nmagnetization dynamics under spin torque using a model\nwhich can incorporate the e\u000bects of TandX. How-\never, the commonly used theoretical model based on the\nLandau-Lifshitz-Gilbert (LLG) equation1,17{19is limited\nat \fxedTdue to the assumption of a \fxed magnetization\nlength (see Appendix A for detailed analysis of the LLG\nmodel). In contrast, the Landau-Lifshitz-Bloch (LLB)\nmodel has been widely used to describe the magneti-\nzation dynamics at elevated T20, where the T-induced\nmagnetization-length change is taken into account by in-\ncluding a longitudinal relaxation term. To date, the LLB\nequations have been implemented for both FMs20and\nFiMs21, and recently the e\u000bect of spin torque in FMs\nhas also been included22. Starting from the atomistic\nLandau-Lifshitz equation, in this work, we extend the\nLLB model to capture the spin-torque e\u000bect in FiMs.\nThe numerical simulation of current-induced switching in\na FiM/heavy-metal (HM) bilayer is then performed, andwe \fnd the modi\fed LLB model can reproduce salient\nexperimental \fndings, such as the magnetization com-\npensation, the reversal of switching direction across the\nmagnetization-compensation temperature ( TMC)23, and\nthe peak in spin torque at TMC24. In addition, the spin-\ntorque-induced sublattice dynamics in FiMs is studied\nand compared to that in AFMs and FMs. The switching\ntrajectory is found to be precession free, and the sublat-\ntices are not always collinear. Finally, the e\u000bect of Xon\nFiM properties is studied.\nII. THE MODIFIED\nLANDAU-LIFSHITZ-BLOCH EQUATION\nAs shown in Fig. 1(a), the device structure we stud-\nied consists of a FiM (Gd X(FeCo) 1X) deposited on top\nof a HM layer. The magnetizations of sublattices are\nmanipulated by the SOT generated by in-plane electrical\ncurrent. The FiM is treated as a two-sublattice model,\ni.e., Gd and FeCo, which is justi\fed by the experimen-\ntal observation that the magnetizations of Fe and Co are\nparallel up to the Curie temperature ( TC)25. The mag-\nnetization dynamics of each sublattice is captured by the\nmodi\fed LLB equation2,3,22,25{30(see Appendix B for the\nderivation)\n_mv=\rv(mv\u0002HMFA\nv )\u0000\u0000v;k(1\u0000(mv\u0001m0;v)\nm2v)mv\n\u0000\u0000v?mv\u0002(mv\u0002m0;v)\nm2v;(1)\nwhere \u0000 v;k= \u0003 v;NB(\u00180;v)=(\u00180;vB0(\u00180;v)) and \u0000 v;?=\n\u0003v;N[\u00180;v=B(\u00180;v)\u00001]=2 are the coe\u000ecients of longitudi-\nnal and transverse relaxation, respectively. The dimen-\nsionless \feld is given by\n\u00180;v=\f\u0016v(HMFA\nv +HI=\u0015v): (2)\nEq. (1) contains two coupled equations for FeCo and Gd\nidenti\fed by the subscript v, which need to be solvedarXiv:1807.02445v1 [cond-mat.mtrl-sci] 6 Jul 20182\nsimultaneously10. The \frst term on the right hand side\ndescribes the magnetization precession around the mean-\nfree \feld\nHMFA\nv =Hext+HA;v+ (J0;v=\u0016v)mv+ (J0;vk=\u0016v)mk;\n(3)\nwhich consists of the external magnetic \feld Hext, the\ncrystalline anisotropy \feld HA;v= (2Dv=\u0016v)mv;zezwith\ncoe\u000ecientDv, and the exchange coupling between sub-\nlattices with coe\u000ecients J0;vandJ0;vk. The damping\ncoe\u000ecient is given by\n\u0003v;N= 2\rv\u0015v=(\f\u0016v); (4)\n\f= 1=(kBT); (5)\nwhere\u0015vis the damping constant, \rvis the gyromag-\nnetic ratio, \u0016vis the magnetic moment, and kBis the\nBoltzmann constant. As previously mentioned, the T-\ninduced magnetization-length change is described by the\nlongitudinal relaxation term using the Brillouin function\nB(\u0018) =coth(\u0018)\u00001=\u0018; (6)\nand the spin-torque e\u000bective \feld HIis given by\nHI=JS\u0016h=(2etFiMjXM S;v\u0000qMS;kj); (7)\nwhere \u0016his the reduced Planck constant, eis the electron\ncharge,tFiMis the thickness of FiM layer, and MSis the\nsaturation magnetization. The spin current density JSis\nformulated as\nJS=\u0012SH\u001b\u0002JC; (8)\nwhere\u0012SHis the spin-Hall angle, \u001bis the polarization\nof spin current, and JCis the charge current. The equi-\nlibrium magnetization m0;vis calculated via the coupled\nCurie-Weiss equation\nm0;v=B(\u00180;v)\u00180;v=\u00180;v: (9)\nSimilar to the LLB in FM22, the e\u000bect of spin torque only\nenters the two relaxation terms. Furthermore, Eq. (1) re-\nduces to Eq. (4) in Ref.21when the spin torque vanishes,\nor to Eq. (7) in Ref.22when FeCo and Gd are not distin-\nguished. The numerical integration of Eq. (1) proceeds\nusing a fourth-order predictor-corrector method10.\nThe parameters used in the simulation are determined\nas follows [see Fig. 1(c)]: First, the Curie-Weiss equation\n[Eq. (9)] for pure Gd and FeCo is solved independently\nand \ft to the experimental M-Tcurves31to determine\nthe exchange coupling coe\u000ecients JGd= 0:98\u000210\u000021J\nandJFeCo = 1:5\u000210\u000021J. Then, the JGdFeCo =\u00007:63\u0002\n10\u000021J is obtained by solving the coupled Curie-Weiss\nequations of FiM. In addition, a su\u000ecient anisotropy is\nused to ensure the perpendicular magnetization. The\n\u0012SHand\u0015are swept with Dto \ft the experimental\nM-HandM-Jloops16, and a good agreement10with\nthe experimental data is obtained with \u0012SH= 0:003732,\n\u0015=0.07, and D= 3:2\u000210\u000026J.\nFiM\nHMt = 3 nmxyz(a)\nSolve Curie -Weiss equation for \nFeCo and Gdindependently\nfit with experimental \n“M vs T”\nJFeCo, JGdfit experimental \n“TMCvs X”Solve Coupled Curie -\nWeiss equation for FiM\nJFeCoGdD,θSH, αSolve LLB equation\nfit experimental MH, MI loop(c)(b)Atomistic LL equation \nwith spin torque for FiMFokker -Planck equation\nMean field model\nFinal form (Eq. 1)FIG. 1. (a) Schematic view of the device structure consist-\ning of a FiM layer deposited on top of a HM. The FiM in\nthis study is perpendicularly magnetized Gd X(FeCo) 1Xal-\nloy, where Gd and FeCo are antiferromagnetically coupled.\nProcedure of (b) equation derivation, (c) model validation\nand parameter determination. The derivation starts from the\nLandau-Lifshitz equation including the spin torque, followed\nby the corresponding Fokker-Planck equation to account for\nthe statistic behavior, and yields the \fnal form after using the\nmean \feld approximation. This model is validated by com-\nparing with the experimental MvsTtrajectory31,TMCvs\nGd concentration ( X), andM-HandM-I loops16.\nIII. DETERMINISTIC SWITCHING INDUCED\nBY THE SPIN-ORBIT TORQUE\nWe \frst study the SOT-induced deterministic switch-\ning in FiM using the LLB model. As shown in Fig. 2, the\nJCapplied along the xdirection generates spin torque\nacting on the FiM layer due to the spin-hall e\u000bect (SHE)\nor the inverse spin galvanic e\u000bect (ISGE)33{35. How-\never, the magnetization cannot be switched vertically\nsince the spin torque aligns the magnetization to yaxis.\nThis is similar to the perpendicular FM switched by\nin-plane current, where an external \feld along the cur-\nrent direction ( HX) is required to achieve determinis-\ntic switching36{39. The switching in FM system can be\nunderstood as follows: The switching direction is deter-\nmined by HXasL= \u0001m\u0002HX, and the spin torque,\n\u0001m=m\u0002(m\u0002HI), should be su\u000ecient to overcome the\nenergy barrier. Therefore, the switching direction will be\nreversed by reversing either HXor current direction37.\nRecently, by applying HX, the current-induced deter-\nministic switching in the FiM/HM bilayer has also been\ndemonstrated4,16. The measured M-Jloop clearly shows\nan opposite switching direction by reversing the current,\nwhereas the e\u000bect of HXhas not been investigated. In\nthis study, we show that the switching direction is also\nreversed under opposite HX[see Fig. 2], which can be\nexplained using the abovementioned two-torque analysis\ntogether with the exchange coupling between sublattices.\nAs shown in Fig. 2(a), the FiM at T= 300 K is FeCo\ndominant. The positive JCandHXswitch mFeCo from\ndown to up, and concurrently, the exchange interaction\nturns mGdfrom up to down. When the HXis reversed,\nmFeCo is switched from up to down [see Fig. 2(b)], re-\nsulting in an opposite M-Jtrajectory. To con\frm the\nunique role of HX, we have veri\fed that the equilibrium3\n-1 0 10.3\n0\n-0.3\nJC(1011A/m2)mZ_FeCo(a)\n-1 0 1\nJC(1011A/m2)(b)\nHXJC\nHXJC\nFIG. 2. Simulated SOT-induced switching in FiM\nGd21(FeCo) 79under (a) HX= 1 mT, (b) HX=\u00001 mT atT\n= 300 K. The HXonly breaks the symmetry for deterministic\nswitching. The reversal of switching direction under opposite\nHXis similarly explained using the theory in perpendicular\nFM.\nmagnetization is not altered when only HXis applied,\nand no switching event is observed when the current is\nswept with Hext= 0 or HY. Therefore, the SOT-induced\nswitching in FiM is determined by the dominant sublat-\ntice, followed by the reversal of the other sublattice via\nexchange interaction, and the HXonly breaks switching\nsymmetry. It is also worth noting that the maximum\nmZin Fig. 2 is around 0.3, which is an evidence of the\nT-induced magnetization-length reduction with mZ= 1\nde\fned atT= 0 K.\nAlthough SOT and HXhave similar e\u000bects in switch-\ning FiM and perpendicular FM, the time evolutions of\nmagnetization are very di\u000berent as shown in Fig. 3,\ni.e., the switching trajectory of FiM is precession free,\nwhereas it is precessional in FM. In the SOT-switched\nFM with initial mZ= 1, both anisotropy \feld and spin\ntorque align the magnetization to the + zdirection for\nmZ>0, resulting in a larger precession term compared\ntomZ<0, where the anisotropy \feld and spin torque\nare opposite. Consequently, more precession occurs when\nmZ>0 [see Fig. 3(b)]. Similarly, the precession-free tra-\njectory in FiM is attributed to the small precession term.\nAs illustrated using the 3D trajectories in Fig. 3(c), mGd\nandmFeCo are switched to opposite directions. Due\nto the strong exchange coupling, many studies assume\nthey are always collinear. However, as the time evolu-\ntion of each sublattice and their relative angle shown in\nFig. 4(a), a maximum deviation of 0.9 degree is observed\natt= 30 ns. This number is similar to a recent report\nfrom Mishra et al.4, where a cant of one degree is es-\ntimated from the strength of exchange \feld. Since the\nexchange coupling between sublattices is very strong ( >\n100 T40,41), even a very small cant deviates the behavior\nof FiM from FM, which might contribute to the di\u000berent\nmagnetization dynamics shown in Figs. 3(c) and 3(d).\nSimilar noncollinearity between sublattices is also pre-\ndicted in AFM42, with the deviation angle determined by\nthe strength of spin torque. To achieve a large-angle non-\ncollinearity in FiM, recent study shows that a magnetic\n\feld over 5 T is required43. By studying the \feld-induced\nswitching in FiM [see Fig. 4(b)], a similar trajectory is\nTime (ns)0 300\n-0.30.3(a)\nmZ_FeComX_FeComY_FeCom\nx\nyz(c)\nmGdmFeCoinitial\nfinal\n(b)\n(d)\nxyzTime (ns)0 10 20\nHXJCFiM\nHXJCFMFIG. 3. Time evolution of the SOT-induced switching (a) in\nFiM Gd 21(FeCo) 79and (b) in perpendicular FM at T= 300\nK. (c) and (d) are the 3D trajectories corresponding to (a)\nand (b) respectively. The dot lines in (c) are the projections\non thex-yplane. All simulations start from the equilibrium\nstate where mZ= 0.3. This reduced value re\rects the T-\ndependent magnetization, where mZ= 1 is de\fned at T= 0\nK.\n(b)\n0\n-0.40.4(a)\nTime (ns)5 25mZ_GdmZ_FeCo\n45179.9\n179.1\n25 45θ\nTime (ns)5\nm\nTime (ns)5 15 250\n-0.30.3\nHXJCFiM\nFIG. 4. (a) Time evolution of mZfor FeCo and Gd sublattices\natT= 300 K, with inset showing the magnetization angle\nbetween mGdandmFeCo. (b) Field-induced switching in\nGd21(FeCo) 79atT= 300 K, which has similar trajectories\nwith the current-induced switching.\nobserved compared to Fig. 3(a), indicating that a large\nspin-torque e\u000bective \feld would be required to get a large\nangle deviation. However, as discussed in the next sec-\ntion, large spin torque aligns the magnetization to the\nspin direction, hence no switching happens.\nIV. EFFECT OF TEMPERATURE AND\nMATERIAL COMPOSITION\nTandXare often tuned in experiments to control the\nproperties of FiM4,9,23,44. By measuring the M-Hloops\nas a function of T,TMCcan be identi\fed where the co-\nercive \feld ( HC) diverges. However, TMCmay not exist\nin another sample with a di\u000berent X16. In this study,\nthe LLB equation is used to investigate two samples, i.e.,4\nmagnetization magnitude vs Tmnet\nT (K)0 100 200 3000\n-0.10.10.2\nmGdmFeCo\nGd21(FeCo)79\nGd23(FeCo )77\nTime (ns)0 15 30\nmZ_FeCo\n0\n-0.80.80\n-0.80.8\nT = 130 K\nT = 70 K(b) (a)\n(c)\nFIG. 5. (a) E\u000bect of Ton the net magnetization of\nGd21(FeCo) 79(blue square) and Gd 23(FeCo) 77(red trian-\ngle) withTMC= 75 K below which Gd is dominant, where\nthe net magnetization is calculated using mnet= (1 \u0000\nX)mFeCo\u0016FeCo +XmGd\u0016Gdwith\u0016FeCo = 2:217\u0016Band\n\u0016Gd= 7:63\u0016B. Time evolution of mZ;FeCo under spin torque\nat (b)T= 130 K and (c) 70 K.\nGd21(FeCo) 79and Gd 23(FeCo) 77, and we show that the\nexistence of TMCis determined by the demagnetization\nspeed and the relative magnitude of mFeCo andmGd.\nAs reported in our recent study10, both mFeCo andmGd\ndecrease with Tand vanish at the same temperature lo-\ncated between TC;FeCo (1043 K) and TC;Gd (292 K). The\ncommon Curie temperature is induced by the strong ex-\nchange coupling which speeds up the demagnetization\nprocess in FeCo but slows down that in Gd. As shown\nin Fig. 5(a), the Gd 21(FeCo) 79shows FeCo dominant at\nall temperatures, whereas a transition from Gd to FeCo\ndominant is observed in the other sample. At low T,\nGd dominates due to the larger magnetic moment. As\nTincreases, mnetreduces and vanishes at TMC= 75\nK because of the faster demagnetization process in Gd.\nAboveTMC,mnetrises until a peak and then reduces to\nzero atTC. Furthermore, we \fnd that the magnetization\ndynamics near TMC[Fig. 5(c)] is similar to the one at\nhigherT[Fig. 5(b)], which can be understood by notic-\ning the gradual change in e\u000bective \felds such as HAand\nHI. It is only at TMCthat a sudden change occurs, and\nthe e\u000bective \felds diverge.\nAs shown in Fig. 6(a), the competition between mFeCo\nandmGdis also manifested in the T-dependent M-H\nloops23. In addition to the reversal of switching direction,\ntheHCreaches maximum at TMCto overcome the en-\nergy barrier ( E=\u0000M\u0001H). WhenTis further increased\n(i.e.,T > T MC), both mFeCo andmGdreduce, result-\ning in smaller exchange and anisotropy \felds [Eq. (3)]\nand hence a lower HC. Furthermore, we \fnd the TMC\nobtained from the M-Hloops is consistent with the equi-\nlibrium state calculation [Fig. 5(a)], which is another ev-\nidence that the LLB model captures FiM dynamics.\nFor practical reasons, Tis not preferred as the con-\ntrol parameter in device applications, whereas Xcan be\ntuned during the deposition process. The change of X\nshows similar results to that observed in the Tdepen-\ndence. AsXis increased, the FiM changes from FeCo to\nGd dominant, resulting in a reversal of both M-Hand\nM-Jloops4,16. Due to the vanishing mnetatXMC, the\nH (mT)-400 -200 200 400-0.60.6mZ_FeCo\n0-0.60.6\n-0.60.6\n-0.60.6\n-0.60.6\n-0.60.610 K\n40 K\n60 K\n90 K\n120 K\n190 KGddominant\nFeCo dominant(a)\nJC(1011A/m2)-8 8-0.60.6mZ_FeCo\n0-0.60.6\n-0.60.6\n-0.60.6\n-0.60.6\n-0.60.6X = 0.15\nX = 0.16\nX = 0.17\nX = 0.24\nX = 0.25\nX = 0.26FeCo dominant\nGddominant(b)Hcrit, Icritvs TFIG. 6. (a) M-Hloops at di\u000berent Tin Gd 23(FeCo) 77with\nHX= 2 mT. The blue dot line denotes the transition from\nGd to FeCo dominant. The switching-direction reversal and\nthe peak in HCobserved in experiments23are qualitatively\nreproduced. (b) M-JCloops at di\u000berent XwithHX= 2\nmT. Three dynamic regions are identi\fed, e.g., for X= 0.25,\nsuccessful switching happens for 5 :7\u00021011A/m2< J C<\n6:4\u00021011A/m2, oscillation region for 6 :4\u00021011A/m28\u00021011A/m2\n(i.e.,maligns to the ydirection).\nspin torque diverges4,44. To show the capability of LLB\nmodel in capturing these e\u000bects, we have simulated the\nX-dependent current-induced switching at T= 300 K.\nAs shown in Fig. 6(b), the switching direction reverses\natXMC= 0.24 which separates FeCo and Gd dominant\nregions. In both regions, mnetis switched from down\nto up under positive current, indicating that the SOT-\ninduced switching is determined by mnet. This is di\u000ber-\nent with the anomalous Hall e\u000bect (AHE), where RAHE\nis determined by mFeCo. In contrast to the magnetic-\n\feld-induced switching in Fig. 6(a), no clear peak of crit-\nical switching current density ( JCrit) is observed, which\nis attributed to the increase of spin torque near XMC. In-\nterestingly, three dynamics regions are identi\fed in our\nsimulatedM-Jloops. According to the sub\fgure of X\n= 0.25 in Fig. 6(b), mFeCo is successfully switched from\nup to down for 5 :7\u00021011A/m2= 2\u0015T\u000eab\u000e(t\u0000t0)=(\r\u00160); (16)\nwhere sis the spin angular momentum, \u0010is the thermal\n\feld with the subscript representing di\u000berent Cartesian\ncomponents (i.e., x,y, andz), andtis the time. The\nthree terms on the right hand side of Eq. (14) represent\nprecession, damping, and spin-torque e\u000bect, respectively.\nThe exchange coupling in the last term of Eq. (15) only\nconsiders the in\ruence of nearest neighbors, and Eq. (16)\nindicates that the sublattice spin is uncorrelated with\nrespect to time and other Cartesian components. The\ndirect simulation using Eq. (14) is known as atomistic\nmodeling2,3,22,25,26, and the information of magnetiza-\ntion dynamics is obtained by summing up all the lattice-\nsite spins. Since the lattice constant is very small (a\nfew angstroms), the atomistic model is limited to very\nsmall devices with diameter below 20 nm26. To simu-\nlate larger devices, a statistical model is developed based\non Eq. (14), resulting in a single equation, i.e., Fokker\nPlanck equation, which captures the spin dynamics as\n@f\n@t+@\n@(N)f\rN\u0002H\u0000\rN\u0002(N\u0002(\u0015H+HI))\n+\r\u0015T\n\u00160[N\u0002(N\u0002@\n@N)]gf= 0;(17)\nwherefis the spin-distribution function, and Nis a vec-\ntor on a sphere with jNj= 1. Then, the spins are trans-\nformed to magnetization through\nm\u0011=Z\nd3NNf(N;t); (18)\nand Eq. (17) becomes\n_m=\r[m\u0002H]\u0000\u0003Nm\u0000\r\u0015< s\u0002[s\u0002H]>: (19)\nHowever, Eq. (19) is di\u000ecult to solve due to the mix-\nture of mands, which can be resolved by applying the\nmean \feld approximation (MFA)21,25, resulting in an ex-\nplicit equation showing as Eq. (1). This process of model\ndevelopment is summarized as a \rowchart in Fig. 1(b).7\n\u0003a0132576@u.nus.edu\nyelelg@nus.edu.sg\n1X. Jiang, L. Gao, J. Z. Sun, and S. S. P. Parkin, Phys.\nRev. Lett. 97, 217202 (2006).\n2I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius,\nH. A. D urr, T. A. Ostler, J. Barker, R. F. L. Evans, R. W.\nChantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk, T. Rasing,\nand A. V. Kimel, Nature 472, 205 (2011).\n3T. A. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell,\nU. Atxitia, O. Chubykalo Fesenko, S. El Moussaoui,\nL. Le Guyader, E. Mengotti, L. J. Heyderman, F. Nolt-\ning, A. Tsukamoto, A. Itoh, D. Afanasiev, B. A. Ivanov,\nA. M. Kalashnikova, K. Vahaplar, J. Mentink, A. Kirilyuk,\nT. Rasing, and A. V. Kimel, Nature Communications 3,\n666 (2012).\n4R. Mishra, J. Yu, X. Qiu, M. Motapothula, T. Venkatesan,\nand H. Yang, Phys. Rev. Lett. 118, 167201 (2017).\n5S.-H. Oh, S. K. Kim, D.-K. Lee, G. Go, K.-J. Kim, T. Ono,\nY. Tserkovnyak, and K.-J. Lee, Phys. Rev. B 96, 100407\n(2017).\n6S. K. Kim, K.-J. Lee, and Y. Tserkovnyak, Phys. Rev. B\n95, 140404 (2017).\n7A. Kamra and W. Belzig, Phys. Rev. Lett. 119, 197201\n(2017).\n8Z. Zhao, M. Jamali, A. K. Smith, and J.-P. Wang, Applied\nPhysics Letters 106, 132404 (2015).\n9K.-J. Kim, S. K. Kim, Y. Hirata, S.-H. Oh, T. Tono,\nD.-H. Kim, T. Okuno, W. S. Ham, S. Kim, G. Go,\nY. Tserkovnyak, A. Tsukamoto, T. Moriyama, K.-J. Lee,\nand T. Ono, Nature Materials 16, 1187 (2017).\n10Z. Zhu, X. Fong, and G. Liang, Phys. Rev. B 97, 184410\n(2018).\n11J. Hohlfeld, T. Gerrits, M. Bilderbeek, T. Rasing,\nH. Awano, and N. Ohta, Phys. Rev. B 65, 012413 (2001).\n12C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk,\nA. Tsukamoto, A. Itoh, and T. Rasing, Phys. Rev. Lett.\n99, 047601 (2007).\n13K. Vahaplar, A. M. Kalashnikova, A. V. Kimel, S. Gerlach,\nD. Hinzke, U. Nowak, R. Chantrell, A. Tsukamoto, A. Itoh,\nA. Kirilyuk, and T. Rasing, Phys. Rev. B 85, 104402\n(2012).\n14A. R. Khorsand, M. Savoini, A. Kirilyuk, A. V. Kimel,\nA. Tsukamoto, A. Itoh, and T. Rasing, Phys. Rev. Lett.\n108, 127205 (2012).\n15J. Barker, U. Atxitia, T. A. Ostler, O. Hovorka,\nO. Chubykalo-Fesenko, and R. W. Chantrell, Scienti\fc\nReports 3, 3262 (2013).\n16N. Roschewsky, T. Matsumura, S. Cheema, F. Hellman,\nT. Kato, S. Iwata, and S. Salahuddin, Applied Physics\nLetters 109, 112403 (2016).\n17C. D. Stanciu, A. V. Kimel, F. Hansteen, A. Tsukamoto,\nA. Itoh, A. Kirilyuk, and T. Rasing, Phys. Rev. B 73,\n220402 (2006).\n18M. Binder, A. Weber, O. Mosendz, G. Woltersdorf,\nM. Izquierdo, I. Neudecker, J. R. Dahn, T. D. Hatchard,\nJ.-U. Thiele, C. H. Back, and M. R. Scheinfein, Phys. Rev.\nB74, 134404 (2006).\n19H. Oezelt, A. Kovacs, F. Reichel, J. Fischbacher, S. Bance,\nM. Gusenbauer, C. Schubert, M. Albrecht, and T. Schre\r,\nJournal of Magnetism and Magnetic Materials 381, 28\n(2015).20D. A. Garanin, Phys. Rev. B 55, 3050 (1997).\n21U. Atxitia, P. Nieves, and O. Chubykalo-Fesenko, Phys.\nRev. B 86, 104414 (2012).\n22P. M. Haney and M. D. Stiles, Phys. Rev. B 80, 094418\n(2009).\n23T. Okuno, K.-J. Kim, T. Tono, S. Kim, T. Moriyama,\nH. Yoshikawa, A. Tsukamoto, and T. Ono, Applied\nPhysics Express 9, 073001 (2016).\n24W. S. Ham, S. Kim, D.-H. Kim, K.-J. Kim, T. Okuno,\nH. Yoshikawa, A. Tsukamoto, T. Moriyama, and T. Ono,\nApplied Physics Letters 110, 242405 (2017).\n25T. A. Ostler, R. F. L. Evans, R. W. Chantrell, U. Atxi-\ntia, O. Chubykalo-Fesenko, I. Radu, R. Abrudan, F. Radu,\nA. Tsukamoto, A. Itoh, A. Kirilyuk, T. Rasing, and\nA. Kimel, Phys. Rev. B 84, 024407 (2011).\n26N. Ulrich, \\Classical spin models,\" in Handbook of Mag-\nnetism and Advanced Magnetic Materials (American Can-\ncer Society, 2007).\n27F. Schlickeiser, U. Atxitia, S. Wienholdt, D. Hinzke,\nO. Chubykalo-Fesenko, and U. Nowak, Phys. Rev. B 86,\n214416 (2012).\n28X. Jiao, Z. Zhang, and Y. Liu, SPIN 06, 1650003 (2016).\n29O. Chubykalo-Fesenko, U. Nowak, R. W. Chantrell, and\nD. Garanin, Phys. Rev. B 74, 094436 (2006).\n30R. F. L. Evans, W. J. Fan, P. Chureemart, T. A. Ostler,\nM. O. A. Ellis, and R. W. Chantrell, Journal of Physics:\nCondensed Matter 26, 103202 (2014).\n31H. E. Nigh, S. Legvold, and F. H. Spedding, Phys. Rev.\n132, 1092 (1963).\n32M. Morota, Y. Niimi, K. Ohnishi, D. H. Wei, T. Tanaka,\nH. Kontani, T. Kimura, and Y. Otani, Phys. Rev. B 83,\n174405 (2011).\n33I. Mihai Miron, G. Gaudin, S. Au\u000bret, B. Rodmacq,\nA. Schuhl, S. Pizzini, J. Vogel, and P. Gambardella, Nat\nMater 9, 230 (2010).\n34A. Manchon and S. Zhang, Phys. Rev. B 79, 094422 (2009).\n35H. Kurebayashi, J. Sinova, D. Fang, A. C. Irvine, T. D.\nSkinner, J. Wunderlich, V. Nov\u0013 ak, R. P. Campion, B. L.\nGallagher, E. K. Vehstedt, L. P. Z^ arbo, K. V\u0013 yborn\u0013 y, A. J.\nFerguson, and T. Jungwirth, Nature Nanotechnology 9,\n211 (2014).\n36I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V.\nCostache, S. Au\u000bret, S. Bandiera, B. Rodmacq, A. Schuhl,\nand P. Gambardella, Nature 476, 189 (2011).\n37L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and\nR. A. Buhrman, Phys. Rev. Lett. 109, 096602 (2012).\n38K. Garello, C. O. Avci, I. M. Miron, M. Baumgartner,\nA. Ghosh, S. Au\u000bret, O. Boulle, G. Gaudin, and P. Gam-\nbardella, Applied Physics Letters 105, 212402 (2014).\n39S. Fukami, T. Anekawa, C. Zhang, and H. Ohno, Nature\nNanotechnology 11, 621 (2016).\n40C. Kittel, Phys. Rev. 82, 565 (1951).\n41R. Khymyn, I. Lisenkov, V. Tiberkevich, B. A. Ivanov,\nand A. Slavin, Scienti\fc Reports 7, 43705 (2017).\n42E. V. Gomonay and V. M. Loktev, Low Temperature\nPhysics 40, 17 (2014).\n43J. Becker, A. Tsukamoto, A. Kirilyuk, J. C. Maan, T. Ras-\ning, P. C. M. Christianen, and A. V. Kimel, Phys. Rev.\nLett. 118, 117203 (2017).\n44N. Roschewsky, C.-H. Lambert, and S. Salahuddin, Phys.\nRev. B 96, 064406 (2017).8\n45K. Cai, M. Yang, H. Ju, S. Wang, Y. Ji, B. Li, K. W.\nEdmonds, Y. Sheng, B. Zhang, N. Zhang, S. Liu, H. Zheng,\nand K. Wang, Nature Materials 16, 712 (2017).46C. J. Gar\u0013 cia-Cervera, Bol. Soc. Esp. Mat. Apl. 39, 103\n(2007).\n47J. Crangle and G. M. Goodman, Proc. R. Soc. London,\nSer. A 321, 477 (1971).\n48T. Devolder, C. Chappert, J. A. Katine, M. J. Carey, and\nK. Ito, Phys. Rev. B 75, 064402 (2007)." }, { "title": "1901.05073v1.SiC_YiG_X_band_quantum_sensor_for_sensitive_surface_paramagnetic_resonance_applied_to_chemistry__biology__physics.pdf", "content": "SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n1 \n \n \n \nSiC-YiG X band quantum sensor for \n \nSensitive Surface Paramagnetic Resonance \n \napplied to chemistry, biology, physics . \n \n \n \nJérôme TRIBOLLET \n \nInstitut de Chimie de Strasbourg, Strasbourg University, UMR 7177 (CNRS -UDS), \n4 rue Blaise Pascal, CS 90032, F -67081 Strasbourg Cedex, France \nE-mail : tribollet@unistra.fr \n \n \n \nABSTRACT \nHere I pr esent the SiC-YiG Quantum Sensor, allowing electron paramagnetic resonance (EPR) \nstudies of monolayer or few nano meter s thick chemical, biological or physical samples \nlocated on the sensor surface . It contains two parts , a 4H-SiC substrate with many \nparamagnetic silicon vacancies (V2) located below its surface, and YIG ferrimagnetic \nnanostripes. S pins sensing properties are based on optically detected double electron -\nelectron spi n resonance under the strong magnetic field gradient of nanostripes. Here I \ndescribe fabrication, magnetic, optical and spins sensing properties of th is sensor. I show \nthat the target spins sensitivity is at least five order s of magnitude larger than the o ne of \nstandard X band EPR spectrometer, for which it constitutes , combined with a fiber bundle, a \npowerful upgrade for sensitive surface EPR . This sensor can determine the target spins \nplanes EPR spectrum, their positions with a nanoscale precision of +/ - 1 nm , and their 2D \nconcentration down to 1/(2 0nm)2. \n \n \n \n \n \n \n \n \n SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n2 \n \n Electron paramagnetic resonance1 (EPR) investigation of electron spins localized \ninside, at surfaces, or at interfaces of ultra thin films is highly relevant . In the field s of \nphotovoltaic2 and photochemistry3, EPR is useful to study the spins of photo -created \nelectron -hole pair s, their dissociation, and their eventual transport or chemical reaction \noccurring at some relevant interface . In opto -electronics with 2D semiconductors4, spins of \ndefects limiting device performance can be identified and quantified by EPR . In magnetic \ndata storage science5 and in spin-based quantum comput ing science using molecules6 \ngrafted, tethered, encapsulated or physisorbed on a so lid substrate , it is relevant to study by \nEPR the magnetic properties of those molecules , always modified by their interaction with \nthe substrate7. In solid supp orted heterogeneous catalysis , it is relevant to study spins \ninvolved in catalytic reactions , using EPR8 and eventually spin trapping methods9. In \nstructural biology , it is relevant to study by EPR spin labeled proteins10,11 introduced in \npolymer suppo rted or tethered lipid bilayers membranes12,13. In the context of the \ndevelopment of new theranosti c agents for nano medi cine, it is relevant to study ligand -\nprotein molecular recognition events occurring on surfaces by EPR, using for example, \nbifunctional spin labels14. As various nanotechnologies now allow to produce nanoscale \nthickness samples , one ne eds to perform sensitive Surface EPR (S -EPR). However, \ncommercial EPR spectrometers have not enough sensitivity15 for EPR study of those few \nmonolayers thick ultra -thin films , particularly when target spins are diluted and when \nsamples stacking is not poss ible. \n Home -made EPR experimental setups have been d eveloped recently , in the context \nof quantum sensors16-20 and quantum computers , reaching single spin sensitivity by \noptically17,18,21, electrically22 or mechanically23 detected EPR . Some of them achiev ed the \nnanoscale resolution imaging , when combined with magnetic devices moving over \nsurfaces24,25. Other recent advances in the field of inductively detected EPR have also \nconsiderably improved sensitivity, but at the price of operating home -made microwav e \ndevices at unco nventional millik elvin temperatures26. Thus, c learly, there is today a gap \nbetween performance s of standard X band EPR spectrometers already used worldwide by \nmost of chemists, biologists and physicists , and the ones of the bests unconvent ional EPR \nsetups found in just few laboratories worldwide . \n Here I present the theory of a new Optically Detected M agnetic Resonance (ODMR) \nbased electron spins Quantum Sensor, allowing to study target electron spins of ultra thin \nparamagnetic samples located on the sensor surface . It has nanoscale resolution in one \ndimension, a high sensitivity due to spins ensemble ODMR , and importantly , is designed as \nan upgrade of standard X band pulsed EPR spectrometers . The design of the magnetic \nproperties of the sen sor is inspired from the one s of the hybrid paramagnetic -ferromagnetic \nquantum computer device27 I previously proposed . However, here, it is adapted to \nconstrain ts of standard X band (10 GHz, 0.35 T, 5 mm sample access ) pulsed EPR resonators \nand spectromet ers and thus to fiber bundle based ODMR28,29. The quantum sensor contains \ntwo parts. The first is a 4H-SiC semiconductor substrate containing, just below its surface, \nisolated negatively charged silicon vacancies (V2) used as quantum coherent ODMR spin \nprobes20,21,30 ,31,45. The second part is an ensemble of ferrimagnetic YIG (Yttrium Iron Garnet) \nnanostripes32 having narrow sp in wave resonances at X band . A fixed spacer fabricated on \nedges adjust the relative distance between the two parts. Next, I present the fabrication \nmethodology, magnetic and optical properties , and finally spins sensing properties , based on \nPELDOR spectroscopy1,10,11, 18,33, of this SiC-YiG quantum sensor. SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n3 \n The quantum sensor device proposed can be obtained by fabricating its two parts \nseparately and then integrat ing them (fig.1 a, b). As said in introduction , the first par t of the \nquantum sensor is a 4H -SiC semiconductor sample , in which silicon vacancies spin \nprobes20,21,30,31 ,45 called V 2 are created just below the 4H -SiC surface, and on which the ultra -\nthin paramagnetic film of i nterest will have to be deposit ed, anchored or self assembled \n(fig.1 ). This is necessary because the spins sensing principle is related to the many long range \ndipolar coupling s that exist between a given singl e V 2 spin probe and the many neighbor \ntarget spins (fig.1c ), those couplings affecting the spin coherence time of V 2 spins probes \nand being revealed by PELDOR spectroscopy1,10,11,18,33. The 4H -SiC sample can be a 4H -SiC \nsubstrate terminated on one side by an isotopica lly purified 4H -SiC grown layer, having no \nnuclear spins21 and a very low residual n type doping (< 1014 cm-3) 21. However, a \ncommercially available 4H -SiC substrate with low n doping and a natural low amount of non-\nzero nuclear spins is also a good starting point . \n \nfigure 1 : a/ two parts of the Q uantum Sensor : the paramagnetic 4HSiC one , with V2 spins on front side of \nthe truncated cone shape island (45°), and a cone shaped dip (45°) on back side; and the ferrimagnetic one, \nwith many identi cal YIG nanostripes on GGG substrate (only one stripe shown here for clarity, thus not at \nscale) . Also shown on b/ , their integration by a spacer (not at scale) and introduction in a standard pulsed \nEPR spectrometer microwave cavity, as well as the fiber b undle and the GRIN microlens (yellow) used for \nfiber bundle based ODMR. b/: Zoom showing the many dipolar couplings (dark lines) existing between V2 \nspins probes in 4H -SiC and target spins in the sample, used for quantum sensing by OD PELDOR spectroscopy. \nMolecular target spins and V2 probe spins a re here separated by a capping layer of few nanometers. W eff \nindicate the width along z direction over which the dipolar magnetic field produced by a nearby YIG \nnanostripe can be considered as homogeneous. dx is t he distance between the plane of V2 spins and t he \nplane of target molecular spins considered here. d1+d2=dx. Orders of magnitude: C 2D,V2= 1/ (30nm)2 et \nC2D,Target= 1/ (5nm)2, dx=10 nm, et d2 =2nm, d1=8 nm, weff =60nm for a nearby YIG nanostripe \n(T=100nm/W=500nm ), whose center is located at a distance x opt=150 nm here from the V2 spin s plane. \n The fabrication process of silicon vacancies V2 spins probes in 4H -SiC that I propose \nhere is described on top of fig.2 . It is based on an implantation -etching approach , combined \nwith SiC sculpting , in order to define the appropriate photonic structure for the optical \nexcitation and detection of V 2 spins probes . After cleaning of th e 4H -SiC surface, 5 nm of \nsacrificial SiO 2 are fabricated on the surface of the 4H -SiC substr ate (thickness of 400 µm) . \nThose 5 nm of Si O2 can be obtained, either by slow oxidation of the 4H -SiC surface34 into \nSiO 2 at around 11 50 °C , or by a lower temperature thin film d eposition method like by \nPECVD35 or atomic layer deposition ( ALD)36 at 150°C . High temperature oxidation should SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n4 \n advantageously remove residual V 2 silicon vacancies initially present in the 3D bulk of the \n4H-SiC sample, as V 2 vacancies are annealed out31 at around 700°C . Then, 20 nm of a \nstopping sacrificial layer of zinc oxide (ZnO ) are deposited on top of SiO 2/4H-SiC, by \nsputtering or by ALD. Then 22 keV As+ ions are implanted in this tri-layer sample at a dose \ncomprised between 1.6 1012 cm-2 and 1.6 1013 cm-2. The target dose here is around 8.3 \n1012cm-2, which corresponds, accordin g to SRIM simulations (see SI), to a 2D effective \nconcentration of As+ ions in the first 2 nm of 4H -SiC of C 2D, As+ = 1/(32nm )2. SRIM simulation s \nalso indicate that the concentration of As+ ions rapidly decay with depth in 4H -SiC and is \nalmost zero after t he first 10 nm of 4H -SiC. SRIM simulations also indicate that such \nimplantation of As+ ions produce 1.3 silicon vacancy per As+ ion in those firs t 2 nm of 4H -SiC. \nOne can thus consider that we obtain a 2D effective concentration of silicon vacancies V 2 in \nthe first 2 nm of 4H -SiC of C 2D, eff, V2 = 1/(32nm )2. This concentration rapidly decays to zero in \nthe next few nanometers in 4H -SiC. Then, 4H -SiC micro -sculpting is performed either by \ndiamond machining37,38, by laser ablation39, by FIB40 or by another mi cromachining method41. \nThe aim is to produce , on front side , a truncated cone shape island with V 2 spins on top, and \non back side , a cone shape dip (cone edge angle of 45° in both cases), both cones sharing the \nsame symmetry axis and having an optical qual ity surface roughness (fig.2 top) . Then , ZnO is \netched by HC l, and SiO 2 is etched by HF42. This leads to a sculpted sample with shallow \nsilicon vacancies created mainly 2 nm below the surface of the 4H-SiC truncated cone shape \nisland . A post implantation -sculpting -etching annealing , at a temperature inferior to 600-\n700°C, can eventually be performed to remove some unwanted created defects . Then, a \ntreatment passivate s the truncated cone shape 4H-SiC island surface, like a H+N plasma \ntreatment43 at 400°C, reducing its surface de nsity of state to 6 1010 cm-2. Then, eventually \n(not shown on fig.2) , a few nm capping layer , easy to functio nalize , can be deposited on th is \npassivated 4H -SiC surface , for example using ALD of silicon oxide at low temperature36. Then, \na spacer of appropriate thickness , 200 nm here , for example a ring shape spacer made of \nsilicon oxide, is fabricated by standard lithography and deposition , on the edge s of th e top \nsurface of the 4H -SiC or 4H -SiC/SiO 2 island , under which the V 2 spins prob es were created . \nThe diameter of this top 4H -SiC island surface is around 900 µm. The spacer will allow the \nintegration of the two parts of the quantum sensor device by contacting them (fig.1 b). \nFinally, the few monolayers paramagnetic film of i nterest can be created on top of the \nsensor surface . It is either chemically anchored or physically adsorbed on the sensor surface , \neventually pre -functionalized. Note also that it is possible to first deposit a nanoscale \nthickness solid thin film on the s ensor surfa ce and then to fabricate a spacer on it, with the \nappropriate thickness. \n The fabrication process of the YIG ferrimagnetic nanostripes array on the GGG \n(Gadolinium Gallium Garn et) substrate, necessary for the second part of this quantum \nsensor (fig.2 botto m), follows process es recently published32,44. Tho se process es were \nsuccessful in producing YIG nanostructured thin films with narrow spin wave resonances at X \nband32,44. Shortly, t hose process es use a reactive magnetron sputtering system operating at \nroom temperature with a YIG target. The deposition has to be done through a mask \nfabricated on GG G, obtained by electron beam lithography (fig.2 bottom ). After the YIG \ndeposition and mask removal , a ther mal treatment at around 750-800°C under air flow or \noxyg en atmosphere, during around 1 or 2 hours , has to be performed32,44. SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n5 \n \n \nfigure 2 : Fabrication of the quantum sensor device : 4H -SiC part (top) and YIG/GGG part (bottom) ; see text \nfor details on the various successive fabrication processes . When possible , and if it is ad vantageous, the \norder of some processes can be modified , as long as the key targeted quantum sensor properties are \nconserved. \n Quantum sensing16-20,24,25 by optically detected17,18,21 PELDOR spectroscopy1,10,11,18,33 \n(fig.5) is only possible if the V 2 spins probes created and coherently manipulated at the \nmicrowave probe frequency fs are sufficiently quantum coherent intrinsically, that is without \nany nearby target spin bath, in order to be able to feel the added spin decoherence1,17,18,24 \nproduced by the spin bath of the sample of study , when it is driven at the microwave pump \nfreque ncy fp (fig.5 ). Let us discuss firstly the electron spin coherence time expected for the \nspin S=3/2 of a 4HSiC silicon vacancy (V 2) 21,30,31 ,45 created by this fabrication process few \nnanometers below the surface. Nuclear spin bath spectral diffusion21 is small in 4HSiC which \ncontains very few non -zero nuclear spins, and it can be elimi nated by isotopic purification . \nBulk electron spin bath spectral diffusion is sm all in lightly n -doped 4HSiC and can be \nreduced by chemical purification and doping control21. Spin-lattice relaxation should be quite \ninefficient for V 2 spins probe s, in view of the very long spin coherence time of 100 µs \nobserved already at room te mperat ure for bulk V 2 spins probes21,30,31. Spin decoherence \ninduced by the residual paramagnetic states present at the 4H -SiC passivated surface is \nnegligi ble for most V 2 spin probes , due to the low 2D residual defect concent ration after \npassivation43 (6.1010 cm-2). Thus, the dominant intrinsic decoherence process for V 2 spins \nprobes in this quantum sensor device is expected to be instantaneous diffusion1 in 2D , \noccu rring among the V 2 spins probe s having the same resonant magnetic field , at fixed \nmicrowave probe frequency and under the strong dipolar magnetic field gradient produced SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n6 \n by the YIG nanostripes . Note that YIG is fully saturated at X band because its saturation \nfield32,44 is Bsat=1700 G and the external B0 field applied for EPR is around 3500 G . \n \nfigure 3 : YIG nanostripes magnetic properties assuming the following d imensions , width W=500 nm , \nthickness T=100 nm, length L=100µm , and Bsat=1700 G . a/ and b/ Electron spin resonance spectrum at X (9.7 \nHz) and Q (34 GHz) band respectively, showing, in b lue, the YIG nanostripes spin wave resonances and, in red, \nthe shifted paramagnetic resonance of reference g=2.00 electron spin s, placed at x opt=150 nm above the YIG \nnanostripe center (x=0) . Paramagnetic and f errimagnetic resonances have linewidth of 1 G h ere. c/ One \ndimensional eigenenergies of the spin waves along z axis (horizontal lines) represented on top of the \ninhomogeneous effective confining potential inside a YIG nanostripe saturated along its width (here \nz*=300+z ) ; z=0 corresponds to the center of the stripe . d/ z component of the dipolar magnetic field of the \nYIG nanostripe as a function of x (black), as well as its gradient along x (red) multiplied here by 100 for \nclari ty. e/ and f/ Total effective Zeeman splitting at X band (dot line), expres sed in Gauss (thus divided by (g \nµB), assuming g=2.00 ), as well as its two contributions: the one of Bdz to first order in blue, and the one of \nBdx in red to second order, as produced by the YIG nanostripe , respectively at xopt (e/) and at xopt -10 nm \n(f/), and both plotted versus z , to show the lateral homogeneity of this effective Zeeman splitting . \n \n The figure 3 summarizes the static and dynamic magnetic properties of the YIG \nnanostripes. The fig. 3d shows that the maximum magnetic field gradient in th e x direction, \nperpendicular to the GGG and 4HSiC surfaces, is of around 0.5 G/nm and is obtained at a \ndistance x opt=150 nm from the center of a given YIG nanostripe. That is why the spacer ha s \nto hav e a thickness of xopt + T/2 =200 nm, such that the V 2 spins probes feel the maximum \nmagnetic field gradient. The magnetic field gradient produced by such a YIG nanostripe is \nnot rigorously one dimensional along x. However , as I previously explained in the context of \nquantum computing27, locally , around x opt = 150 nm here , and laterally at z=0 +/ - 30 nm \nalong z, detailed calculations clearly show (fig. 3e) that in this portion of plane above each \nYIG nanostr ipe, the dipolar magnetic field can be considered as laterally homogeneous with \na precision of 0.1 G. Even in the portion of plane located at around x opt - 10 nm, and laterally \nat z=0 +/ - 30 nm along z, which is a possible position where target spins could be found, the \ndipolar magnetic field can be considered as laterally homogeneous with a precision of 0.3 G SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n7 \n (fig. 3f) . As the V 2 spins probes in 4HSiC have a narrow linewidth21,30,31 ,45,46 of less than 1 G, \nwith a gradient here of 0.5 G/nm, one can thus consider that all the V 2 spins probes located \nbetween xopt and xopt-2nm (fig 1 c), just below the 4HSiC surface , and with z=0 +/ - 30 nm \nalong z (weff=60 nm) , have the same resonant magnetic field with a precision of around 1 G. \nAs their 2D concentration obtained by fabrication is 1/(32nm)2, their decoherence time \nassociated to instantaneous diffusion in 2D is nume rically calculated to be T ID,2D= 12.5 µs, \nand i s independent of the temperature . Selective microwave pulses1 can thus excite th is V2 \nspins probes plane , without exciting the other more diluted V2 spins planes located in the \nnext few nanometers of 4HSiC . The V2 plane - target spins plane distance is thus measured \nhere with a precision of around +/- 1nm. \n \nfigure 4 : Some optical properties of the quantum sensor described here: a/ ODMR setup: fibers bundle (6+1, \nin blue), GRIN lens (NA=0.5, 0.25 pitch, diam eter: 500 µm, in yellow ) for collimation after the central fiber, \nEPR tube (in gray), and 4H -SiC sculpted sam ple (edges in red, cone angles are 45°); all are inserted inside a \nmicrowave resonator like the MD5 flexline resonator (the YIG part of the sensor , supporting the SiC one, is \nnot shown here for clarity); also shown on a/, static B0z and microwave magnetic field B 1x(t), some near \nsurface V 2 electric dipoles aligned along the c axis of 4H -SiC (in violet, maximum emission along the z axis , \northogonal to the c axis), and some relevant optical rays for geometric optics investigation of the excitation \nand collec tion efficiencies of this new ODMR based setup for quantum sensing. Blue ray is an optical \npumping ray with many TIR on SiC faces. Black rays are also optical pumping rays, but TIR are not shown for \nclarity. Violet rays are photoluminescence rays emitted a t 10° with respect to the horizontal and they are still \ncollected by TIR in lateral fibers (NA=0.44, diameter: 500µm). See also zoom in SI. b/ sec tion view of the fiber \nbundle just above the SiC sample. c/Negatively charged silicon vacancy V 2 energy level scheme, explaining \nthe optical readout cycle and the optical pumping cycle. Level names31,4 5,46: 1: (Ground State, S=3/2, M sz= -\n3/2 ( or +3/2 )), 2: (Excited State) , 3: (Meta -stable excited state) , 4: (Ground State, S=3/2, M sz= -1/2 ( or +1/2 )), \nk12 is lase r induced optical absorption/emission rate, k 21 is photoluminescence rate, k 23=k32=kISC is the \nintersystem crossing rate, k 34 is a non-radiative relaxation rate. d/ and e/ : Numerical simulations of \npopulations, based on rate equations, showing the optical pumping31,4 5,46 time necessary to saturate the \npopulation of V 2 spins in the - 1/2 states (green curve) to its maximum value of 0.5 (Note: one can also show \nthat under such OP, the population of V 2 spins in the + 1/2 state also saturates to 0.5, using a similar energy \nlevel scheme and OP/OD cycles). In d/, k23=k32=1 /(17 ns) at 300K31,4 5,46, and in e/, k23=k32=1/(1700 ns) \nassumed at 5K, and for bo th, k21=1/(6ns), k 34=1/(107 ns), k 12sat=2.6 ns-1. Populations shown: N 1 in black, N 2 in \nred, N 3 in blue, N 4 in green. One finds an optical pumping time of around 20 µs at 5K, and 2 µs at 300K, with \nthose parameters. \n SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n8 \n \n It must be also noted here that microwa ve driving of any spin wave resonance of the \nYiG nanostripes of the quantum sensor, during the ODPELDOR sequence used for quantum \nsensing, would add unwanted decoherence27 to V 2 spins probes. That is why the \nferrimagn etic insulating YIG nanostripes were ca refully designed here such that there is no \nspectral overlap between their confined spin wave resonances27 (fig. 3 a, b, c), which are \nnarrow in YIG32,44, and the shifted paramagnetic resonances of the V 2 spins probes (fig. 3a, b). \nNote also that according to my previous theoretical calculation27, thermal fluctuations of Y iG \ndo not contribute to decoherence of V 2 spin probes, due to the reduce d saturation \nmagnetization of Yi G compared to the one of Permalloy previously considered in the context \nof quantum co mputing27. Note also that, as instantaneous diffusion is temperature \nindependent and as Y iG is still ferrimagnetic at room temperature, this hybrid Si C-YiG \nquantum sensor can be used in principle between 4K and 300K. \n \n The ODMR at X band of the ensemble o f V 2 spins probes used for sensitive quantum \nsensing , is based on efficient optical pumping21,30,31,4 5,46 (fig. 4 a,c,d,e ), as well as on the \nefficient collection of V 2 spins probes photoluminescence21,30,31,4 5,46 (fig. 4 a,b), by means of a \nfiber bundle28,29, a small GRIN microlens (fig . 1a,b and fig. 4 a,b ), and the man y total internal \nreflexion19 (TIR) occuring both in the sculpted 4HSiC sample (n=2.6) and in the optical fibers \n(fig 4 a and see also SI) . All components of this ODMR setup can be introduc ed inside \nstandard X band pulsed EPR microwave resonator1,29 allowing PELDOR spectroscopy, like the \nMD5 flexline resonat or47, which accept EPR tubes with external diameter up to 5 mm . \n One can show that the photoluminescence signal Spl, integrated during T by the \nphotodetector, in the ODPLEDOR sequence (fig. 5a ), is given by ( see SI ): Spl = S0.(1-f) , with \nS0 = pex.pcoll.pdet.(T/ԎV2).(N V2/8) and f, a function that depends on the parameters: 2.t1, 2.t2, \nTid,2D, td, C 2D,T, pB(fpump) (see SI for definitions and details). Note that pB(fpump) is equal to 1 \nwhen fpump equal the target spins resonant frequency, and 0, when fpump is far o ff reso nance \nwith the target spin s resonant frequency . In optimal experimental conditions, the Noise N pl \nis dominated by optical shot noise, Npl = (Spl(pB=0))0.5. Thus the \"net signal\" to \"noise\" ratio R \nis given by R=(S pl(pB=1) - Spl(pB=0 )) / Npl . The detailed sensitivity analysis of this quantum \nsensor (see SI) shows, that in optimal exper imental conditions, one could obtain the 200 \nMHz ODPELDOR spectrum shown on fig. 6b (100 points, one point each 2MHz assumed \nhere) in 1.2 s, with a large signal to no ise ratio R=2600. \n The numerically simulated (see SI) spins quantum sensing properties, o btained by \nODPELDOR (fig.5a), are shown on fig. 6. The figure 6a presents the shifted field sweep EPR \nspectrum at 9.7 GHz of V 2 spins probes located at x opt= 150 nm f rom YIG nanostripes (in \ngreen) and of two kinds of target spins S=1 located at x opt-dx=145 nm, that is on the sensor \nsurface (in blue and red, see legend for details), as it could be obtained by direct detected \nEPR, if it would be sensitive enough for Surf ace Paramagnetic Resonance. The edge spin \nwave resonance of Y iG nanostripes having the hig hest resonance field at 9.7 GHz has also \nbeen added to this spectrum (in pink). The shifted EPR line of V 2 at highest field is chosen \nhere for ODPELDOR, which means t hat B 0z is set to this field resonance value, and fs is set to \n9.7 GHz, while f pump is sca nned during ODPELDOR (fig. 5a). The figure 6b shows the resulting \nexpected X band ODPELDOR spectrum versus f pump-fs, scanned over around 200 MHz. The \nfigure 6c indica tes how the normalized ODPELDOR net signal to noise ratio (see SI), given by \nR/R opt= 1-VDeer(td, dx, C 2D,T), depends on 1-VDeer, VDeer being the DEER11 signal, and thus how \nit depends on the relative distance dx between spins probes plane and target spins plane, on SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n9 \n the target spin plane concentration C 2D,T , and on time constant td. Thus clearly, this SiC -YiG \nquantum sensor can determine rapidly the target s pins plane EPR spectrum and its 2D \nconcentration down to 1/(20nm)2, with a sufficiently high net sign al to noise ratio, still \nassuming a V 2 spins probes planar concentration of 1/(32nm)2, and an associated \ninstantaneous diffusion decoherence time in 2D of T ID,2D= 12.5 µs. \n \n \n \n \nfigure 5: a/ X band OD -PELDOR quantum sensing sequence and b/ X band ODMR spin echo decay sequence \nfor characterization of spin coherence time T2 of V 2 spins probes. The spins states -1/2 and +1/2 are prepared \nsimultaneously by optical pumping (laser pulse of 100 µs assumed here). The microwave probe frequency fs, \nand static fie ld B0z, are adjusted to obtain the paramagnetic resonance at this frequency fs with the chos en \noptically pumped EPR transition of V 2 probes spins, either (-3/2 < --> -1/2), or (+1/2 < --> +3/2 ). Both a/ and b/ \ntime resolved ODMR experiments corresponds nearl y to standard PELDOR and Echo Decay experiment s1, but \nthey start after optical pumping and t hey are complemented by a last + /-Pi/2 pulse in order to transform \ntransverse magnetization Mx (tSRT - T - twait), into populations of V 2 spins , which have differen t spin \ndependent photoluminescence and relaxation properties under laser excitation . This allow s the final optical \ndetection of EPR, the so-called spins ensemble ODMR , by means for example, of a gated Photomultiplier \ntube (PMT) . As a first approximation here, and to better understand the hybrid optical -microwave pulses \nsequences , spins states -1/2 and +1/2 are assumed Dark states, while spins states -3/2 and +3/2 are assumed \nBright photo -luminescent states45,46. \n Now I compare the sensitivity of this Si C-YiG fiber bundle based ODMR quantum \nsensor with other setups. Firstly, it must be noted that the same ODPELDOR spectrum as the \none of fig.6b could be obtained also in 1.2 s with a quantum sensor having a single V 2 spin \nprobe, assuming identical experimen tal parameters, but at the price of a reduced net signal \nto noise ratio of only R=2 (see SI) . This new spin ensemble quantum sensor48 is thus 1000 \ntimes more sensitive than a similar single spin-based quantum sensor . It is thus \nadvantageous in terms of bot h measurement time and sensitivity. Of course, ensemble \nmeasurements imply an additional statistical averaging of target spins plane properties, \nwhich is not present in single spin probe measurements, but such stati stics is often a \nrelevant information, li ke in biology10,11 and in realistic solid state devices5,6. Also, this spins \nensemble quantum sensor has a nanoscale spatial resolution in 1D due to the static gradient SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n10 \n used, but no scanning and thus no 3D imaging capabilities, contrary to some scanning s ingle \nspin sensors. Thus, those two kinds of quantum sensors are quite complementary research\n \n \n \n \nfigure 6 : Spins sensing properties of the quantum sensor. a/ The theoretical shifted field sweep EPR \nspectrum at fs= 9.7 GHz of spins S=3/2 of V2 spins probes (giso=2.0028, uniaxial magnetic anisotropy along c \naxis Dc= +35 MHz, C3V) located at xopt =150 nm (in green) and of two different ensembles of anisotropic \nmolecular nanomagnets with target spins S=1 (S1=1, g iso,1=2.0028,D c,1=20MHz, C3V, and S 2=1, \ngiso,2=2.0028,D c,2=180MHz, C3V) located at xopt -dx=145 nm here , thus on the sensor surface (assuming 3nm of \nSiO2 capping layer ). V2 spins and nanomagnets are assumed here to have their C3V c axis orthogonal to B0z. \nEPR simulation in a/ performed with Easyspin software. b/ ODPELDOR spectrum versus fpump -fs, \nassociated to spectrum a/, assuming B0 z is set equal to the highest EPR resonance of V2 on a/. c/ \nDependence of ODPELDOR normalized net signal to noise ratio (see SI) , R/Roptimum=1 -V, on the relati ve \ndistance dx between spins probes plane and target spins plane (dx= 5 nm (1/) , 10 nm (2/), or 15 nm (3/), \nfrom top to bottom), as well as on the target spin plane concentration (C 2D,Target =1/(d2), with d in nm). Dark \ntrace is for td= 5µs, red trace is for td=3µs, blue trace is for td= 1µs (see fig.5 for definition of td) . \n \ntools. However, this new quantum sensor based on spins probes ensemble , has not only the \nadvantage of being much more sensitive and fast er, but also to be compatible with standard \nX ba nd pulsed EPR spectrometers, such that it should be widely used in a soon future by \nmany researchers , already using standard EPR and who want to improve its performances . \nThe detailed comparison (see SI) of the sensitivity of standard X band direct inducti vely \ndetected EPR (DD -EPR) with the one of this quantum sensor upgrade d EPR (noted here \nQUSU -EPR) , shows that t he sensitivity gain on target spins number is at least of five orders of a/ b/ \nc/ dx = 5 nm \ndx = 10 nm \ndx = 15 nm SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n11 \n magnitude . It thus clearly allows to perform surface EPR using this quan tum sensor \ncombined with a commercial X band pulsed EPR spectromete r and an optical fiber bundle . \nThis quantum sensor upgraded EPR spectroscopy should thus open new research directions , \nlike in the field s of surface chemistry and photovoltaic , in structur al biology and \nnanomedicine , as well as in optoelectronics , spintronics and quantum information \nprocessing . \n As a last remark, one can note that th is theoretical work, as well as the e xperimental \ndevelopment29 of this hybrid SiC -YiG quantum senso r, can be viewed as intermediate steps \ntowards the future development of an intermediate scale hybrid YiG -SiC spins qubits -based \nquantum computer, following the guidelines I previously published27. This not scalable \nquantum computer design could however still be very useful for efficient quantum \nsimulations of new potential molecular drugs49. The advantage s of th is YiG-SiC quantum \ncomputer proposal compared to my previous Permalloy -SiC quantum computer proposal are, \nthe n arrow spin wave resonances of Yi G, the coherent microwave manipulations of SiC spin \nqubits at the standard X band, optical initialization and optical detection of EPR of spins \nqubits ensemble , and probably a high operation temperature for SiC spins qubits, some of \nthem remain ing quantum coherent over hundred microseconds, even at room \ntemperature21,30. \n \n \nREFERENCES : \n1/ Principles of pulse electron paramagnetic resonance (2001) . A. Schweiger and G. Jeschke, \nbook from Oxford University Press, Oxford UK; New York (2001). \n \n2/ Time Resolved EPR study of electron -hole dissociations influenced by alkyl side chains at \nthe photovoltaic polyalkylthiophene:PC BM interface . T. Miura et al., J. Phys. Chem. Lett. \n(2014), 5, p 30. \n \n3/ EPR investigation of photoinduced radical pair formation and decay t o a triplet state in \na carotene -porphyrin -fullerene triad . D. Carbonera et al., J. Am. Chem. Soc. (1998), 120, p \n4398. \n \n4/ Paramagnetic intrinsic defects in polycristalline large -area 2D MoS2 films grown on SiO2 \nby Mo sulfurization . A. Stesmans et al., Na noscale Res Lett. (2017), 12, p 283. \n \n5/ Magnetic memory from site isolated Dy (III) on silica materials . F. Allouche et al., ACS \nCent. Sci. (2017), 3, p 244. \n \n6/ A porphyrin spin qubit and its 2D framework nanosheets . A. Urtizberea et al., Adv. Funct. \nMater. (2018), 28, p 1801695. \n \n7/ Probing magnetic excitations and correlations in single and coupled spin s ystems with \nscanning tunneling spectroscopy. M. ternes, Progress in Surface Science (2017), 92, p 83. \n SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n12 \n 8/ EPR characterization and reactivity of su rface -localized inorganic radicals and radicals’ \nions. M. Chiesa et al., Chem. Rev. (2010), 110, p 1320. \n \n9/ Electron paramagnetic resonance spectroscopy of catalytic surfaces. M. Chiesa et al., \nChem. Rev. (2010), 110, p 1320. \n \n10/ Identifying conformation al changes with site -directed spin labeling. W.L. Hubb el et al., \nNature Structural Biology (2000), 7, p 735. \n \n11/ Direct conversion of EPR dipolar time evolution data to distance distributions. G. \nJeschke et al., Journal of Magnetic Resonance (2002), 155, p.72. \n \n12/ Tethered and polymer supported bilayer lipi d membranes: structure and function. J. \nAndersson et al., Membranes (2016), 6, 30, p 1 -14. \n \n13/ Polymer supported lipid bilayers. I. P. McCabe et al., Open Journal of Biophysics (2013), \n3, p 59. \n \n 14/ A bifunctional spin label for ligand recognition on surfaces . M. A. Hollas et al., Angew. \nChem. Int. Ed. (2017), 56, p 9449. \n \n15/ Electron spin resonance probe based on a 100 µm planar microcoil. G. Boero et al., \nReview of Scientific Instruments (2003), 74, p 4794. \n \n16/ Quantum sensing. C.L. Degen et al., Rev. Mod. Phys. (2017), 89, p 35002. \n \n \n17/ Nanoscale sensing using point defects in single crystal diamond: recent progress on \nnitrogen vacancy center -based sensors. E. Bernardi et al., Crystals (2017 ), 7, p 124. \n \n18/ Sensing external spins with nitrogen -vacancy diamond. B. Grotz et al., New Journal of \nPhysics (2011), 13, p 55004. \n \n19/ Broadband magnetometry and temperature sensing with a light -trapping diamond \nwaveguide. H. Clevenson et al., Nature Physics (2015), 11, p 393. \n \n20/ Magnetic field and temperature sensing with atomic scale spin defects in silicon \ncarbide. H. Kraus et al., Scientific Reports (2014), 4, article number: 5303. \n \n21/ Coherent control of single spins in silicon carbide at room temperature. M. Widmann \net al., Nature Materials (2015), 14, p 164. \n \n22/ Single shot readout of an electron spin in silicon. A. Morello et al., Nature (2010), 467, p \n687. \n \n23/ Single spin detection by magnetic resonance force microscopy. D. Rugar et al., Nature \n(2004), 4 30, p 329. SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n13 \n \n24/ Subnanometer resolution in three -dimensional magnetic resonance imaging of \nindividual dark spins. M. S. Grinolds et al., Nature Nanotechnology (2014), 9, p 279. \n \n25/ Nanoscale spin manipulation with pulsed magnetic gradient fields from a ha rd disc \ndrive writer. S. Bodenstedt et al., Nano Lett. (2018), 18, p 5389. \n \n26/ Reaching the quantum limit of sensitivity in electron spin resonance. A. Bienfait et al., \nNature Nanotechnology (2016), 11, p 253. \n \n27/ Hybrid paramagnetic -ferromagnetic qua ntum computer design based on electron spin \narrays and a ferromagnetic nanostripe. J. Tribollet, Eur. Phys. J. B. (2014), 87: 183. \n \n28/ Implementation of optically detected magnetic resonance spectroscopy in a \ncommercial W -band cylindrica l cavity. G. Jans sen et al., Rev. Sci. Instrum. (2001), 72, p 4295. \n \n29/ First experimental development s towards quantum sensing with a standard X band \nEPR spectrometer and a hybrid paramagnetic -ferrimagnetic quantum sensor device . J. \nTribollet, to appear in 2019 . \n \n30/ Point defects in SiC as a promissing basis for single -defect, single -photon spectroscopy \nwith room temperature controllable quantum states. P.G. Baranov et al. , Materials Science \nForum (2013), 740 -742, p 425. \n \n31/ Optical spectroscopy on silicon vacancy defects in silicon carbide. Franzsiska Fuchs, PhD \nthesis (2017), Wurzburg University. \n \n32/ Epitaxial patterning of nanometer -thick Y 3Fe5O12 films with low magnetic damping . S. \nLi et al., Nanoscale (2016), 8 (issue 1), p 388. \n \n33/ Three p ulse ELDOR theory revisited. K.M. Salikhov et al., Appl. Magn. Reson. (2014), 45, \np 573. \n \n34/ Growth rates of dry thermal oxidation of 4H -silicon carbide. V. Simonka et al., Journal \nof Applied Physics (2016), 120, p 135705. \n \n35/ Influence of PECVD of SiO2 passivation layers on 4H -SiC Schottky rectifiers. S. Nigam et \nal., Electrochem. Solid state Lett. (2003), 6, G4 -G6. \n \n 36/ Low temperature silicon dioxide by thermal atomic layer deposition: investigation of \nmaterial properties. D. Hiller et al., Jo urnal of Applied Physics (2010), 107, p 64314. \n \n37/ The current understanding on the diamond mach ining of silicon carbide. S. Goel, J. \nPhys. D: Appl. Phys. (2014), 47, p 243001. \n \n38/ Brittle -ductile transition during diamond turning of single crystal sil icon carbide. S. \nGoel et al., International Journal of Machine Tools and Manufacture (2013), 65, p 15. SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n14 \n \n39/ Review of laser microscale processing of silicon carbide. B. Pecholt et al., Journal of \nLaser Applications (2011), 23, p 12008. \n \n40/ Solid immer sion lenses for enhancing the optical resolution of thermal and \nelectroluminescence mapping of Ga N-on-SiC transistors. J.W. Pomeroy et al., Journal of \nApplied Physics (2015), 118, p 144501. \n \n41/ Machining processes of silicon carbide: a review. P. Pawa r et al., Rev. Adv. Mater. Sci. \n(2017), 51, p 62. \n \n42/ The effectiveness of HCl and HF cleaning of Si 0.85Ge 0.15 surface. Y. Sun et al., Journal of \nVacuum Science and Technology A (2008), 26, p 1248. \n \n43/ Chemical and electronic passivation of 4H -SiC s urface by hydrogen -nitrogen mixed \nplasma. B. Liu et al., Applied Physics Letters (2014), 104, p 202101. \n \n44/ Patterned growth of crystalline Y 3Fe5O12 nanostructures with engineered magnetic \nshape anisotropy. N. Zhu et al., Applied Physics Letters (2017), 110, p 252401. \n \n45/ Spin and optical properties of silicon vacancies in silicon carbide - a Review. S.A. \nTarasenko et al., Phys. Status Solidi B (2018), 255, p 1700258. \n \n46/ Highly efficient optical pumping of spin defects in silicon carbide for stimu lated \nmicrowave emission. M. Fischer et al., Phys. Rev. Applied (2018), 9, p 54006. \n \n47/ Exploiting the symmetry of the resonator mode to enhance PELDOR sensitivity. E. \nSalvado ri et al., Appl. Magn. Reson. (2015), 46, p 359. \n \n48/ Subpicotesla diamond m agnetometry . T. Wolf et al., Phys . Rev. X (2015), 5, p 41001 . \n \n49/ Hardware -efficient variational quantum eigensolver for small molecules and quantum \nmagnets . A. Kandala et al., Nature (201 7), 549, p 242. \n \n \n \n \n \n \n \n \n \n \n \n \n \n SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n15 \n \nSUPPLEMENT ARY INFORMATIONS \n The number of V 2 spins probes having the same resonant magnetic field placed at \nxopt=150 nm above a given YIG nanostripe (500 nm*100 nm*100 µm) , and within an \neffective width of W eff =60 nm around z=0 (fig. 1b), is estimated to be at least equal to 3000 , \ntaking C2D, V2=1/(32nm)2. Assuming the YIG nanostripes are laterally separated by 5 µm, one \nhas an ensemble of around 500 identical YIG nanostripes over the useful squar e surface of \nthe sensor estimated to be Su= 500 µm*500 µm, taking into account the spac er width. Thus, \none has around 1.5 106 identical V 2 spins probes on the sensor surface which have the same \nresonant magnetic field at fixed microwave frequency, that means under the strong gradient \nproduced by the nanostripes . Note also that the surface S* associated to target spins having \nthe same resonant magnetic field is approximately given by S*= (60 nm*100 µm) * 500 = 0.3 \n10-4 cm2. \n The ODMR at X band of the ensemble of V 2 spins probes used for quantum sensing, \nis based on efficient optical pumping21,30,31,4 5,46 (fig. 4a and 4b), as well as, on efficient \nphotoluminescence collection21,30,31,4 5,46 (fig. 4b and 4c) of the V2 spins probes in the 4H -SIC \nsculpted sample , by means of a fiber bundle containing seven fibers (fig. 1a and 4c) and of a \nsmall GRIN (gradient index) microlens (fig. 1a) , as described in details below . \n The central fiber sends exciting l ight, for example at 780 nm or at 805 nm , along an \noptical axe common to the GRIN microl ens and to the cone shape dip of the 4H -SiC substrate \n(45° is the half angle of the cone) . The GRIN lens, 0.25 pitch plan -plan, allows collimation of \nthe light emerging from the central fiber. Then , by means of a first refraction at the \nair(Helium)/SiC interface and then by means of the many total internal refle xions (TIR) \noccuring inside the SiC substrate (n=2.6) (fig.4b), the geometric configuration of the 4H -SIC \nsculpted sample allow s many optical rays to excite the V2 spins located on the useful sensor \nsurface at the top of the truncated cone shape 4H -SIC isl and. This TIR strategy is inspired \nfrom a previous one adopted for sensors fabrica ted with NV centers in diamond19, but with \nhere a different sample design , difficult to implement w ith diamond technology , because \ndiamond is harder than Si C and diamond has not a single defect axis common to all spins \nprobes , like the V 2 center in 4H -SiC (the c axis of 4H -SiC is the only axis for V 2). This new \ndesign allows both optimization of optical excitation and of photoluminescence collection in \nthe restricted volume of an EPR tube of less than 5 mm in external diameter , as required for \nusing standard X band pulsed EPR resonator and spectrometer . Note that t he oblique \nincidence of the exciting light at the senso r surface (incidence angle of around 29° on sensor \nsurface with this design ), after the first refraction , provides a non zero optical electric field \ncomponent parallel to the c axis and thus allows the efficient V2 electric dipole \nexcitation21,30,31,4 5,46.Here, I also assume that the optical excitation power at 780 nm or at \n805 nm, at the output of the central fiber , is sufficiently high to allow the full saturation of \nthe optical transition , during OD and OP sequences. It was previously shown46 that the \noptical power necessary to obtain saturation values of optical V2 spins pumping is inversely \nproportional to their longitudinal spin -lattice relaxation time T1(T), at the temperature T . As \nT1(T) increases up to several tens of second at 5K46, then less than 1 mW at 780 nm spread \nover a 1mm*1mm square sample is suffic ient at 5K for obtaining such optical pumping \nsaturation. Of course, at room temperature, much more power is required, typically more SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n16 \n than 100 mw46. Thus, from the above considerations, I consider here an optical excitation \nefficiency for V 2 spins located on the useful sensor surface of pex=1. \n The photoluminescence of excited negatively charged silicon vacancies V2 in 4H -SiC is \nemitted at 915 nm at low temperature (zero phonon line21,30,31,4 5,46 at 5K ). The excited V2 \nelectric dipoles, aligned along the c axis of 4H -SiC, emit their photoluminescence \npreferentially in the plane perpendicular to the c axis, which means here, at the horizon tal. \nThe edges at 45° of the truncated 4H-SiC cone shape island thus allow , by one reflexion , to \ndirect most of the V 2 spins probes photoluminescence vertically, towards the six lateral \nfibers, in which it is efficiently propagated by TIR, till the infrared photoluminescence \ndetector . In order to evaluate more quantitatively the collection efficiency of this fiber \nbundle based optical setup , defined as the ratio of the collected optical power over the \nemitted optical power by V 2 dipoles , one can use the classical model of a linear dipole \naligned along the c axis for the V 2 dipole and its emission profile determined using the \nPointing vector expression . Using geometric optics (see fig. 4a) and considering the various \ndimensions of the setup and the relevant refractive index of the materials of the setup \n(nSiC=2.6, n air=1, an d for fibers n glass=1.5 and NA=0.44), one can determin e that almost all rays \nemitted by the V2 dipole s of the useful sensor surface around the horizontal direction at +/- \n10° (= π/18 radians ), can, after relevant reflexion s (TIR) on the 4H -SiC sample surfaces, enter \ninto the lateral optical fibers with a suff iciently small angle such that TIR allows the \npropagation of those rays without loss till the end of the fibers, towards the photodetector. \nConsidering the Pointing vector expression associated to the V 2 dipole in spherical \ncoordinates , one can approximate the collection efficiency pcoll by the ratio between the \nemitted PL and the collected PL, assuming that the PL is collected by the fiber bundle setup \nwhen Ө is comprised between (π/2 - π/18) and (π/2 + π/18). \npcoll is thus given by the formula : \npcoll = ( ꭍ sin3(Ө) dӨ, π/2 - π/18, π/2 + π/18) /( ꭍ sin3(Ө) dӨ, 0, π) \nand thus one finds here pcoll = 0.25 . \n The photodetector can be a near infrared sensitive photomultiplier tube with low \ndark counts, or another low noise in frared photodetect ion setup . Here I assume a standard \ninfrared photodetector efficiency pdet=0.01 . Note also that the bundle is divided, outside the \nstandard EPR cryostat (like the CF935 fro m OXFORD for Bruker EPR resonators ), into a single \nfiber, the central one used for optical excitation, and in to a bundle of the six lateral fibers \ncollecting the photoluminescence, further directed towards the photodetector. \n \n Now let us evaluate the net s ignal to noise ratio R of this ODPELDOR experiment and \nthen the sensitivity of this YiG-SiC fiber bundle -based quantum sensor. Starting from the \nDEER experiment expression1,11,33, directly related to the ODPELDOR experiment shown on \nfig. 5a, and considerin g the optical detection of V 2 spins probes and thus the last additional \nπ/2 microwave pulse, one obtains a photoluminescence signal expression Spl, integrated \nduring T by the photodetector , given by: Spl = S0.(1-f) , with S0 = pex.pcoll.pdet.(T/ԎV2).(N V2/8) \nand f, a function that depends on the parameters: 2.t1, 2.t2, Tid,2D, td, C 2D,T, pB. The function f \nis given by f = exp( -((2.t1 + 2.t2)/ Tid,2D)2/3).( (1-pB) + p B.Vdeer(td, dx, C2D,Target ) ), where Vdeer is \nthe standard DEER signal. It can be numerically computed using the linear approximation \nand shell factorization model11. This model was previously introduced for calculating the \nstandard DEER time domain signal in the case of a three -dimensional distributions of spins. \nHere, this model ha s been adapted to take into account the bidimensional random \ndistribution of the target spins in their well-defined plane, parallel to the SiC substrate SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n17 \n surface . The function p B depends on the frequency detuning between the microwave pump \nfrequency and the target spin resonance frequency at f ixed B 0z. Thus , pB=1 on resonance , \nand p B=0 far off resonance for an appropri ate duration π microwave pulse . The function p B is \ngiven by the us ual probability transition formula describ ing the Rabi oscillation between the \ntwo appropriate spins quantum states under application of a microwave pulse . \n \nIn optimal experimental conditions, the Noise N pl is dominated by the optical shot noise, \ngiven by Npl = (Spl(pB=0))0.5. Thus the \"net signal\" to \"noise\" ratio R is given by the formula \nR=(S pl(pB=1) - Spl(pB=0 )) / Npl . Thus, introducing Ropt, the optimal signal to noise ratio, R is \ngiven, in the general case, by: R=R opt*(1- VDeer(td, dx, C2D,T)), with Ropt given by the formula: \nRopt = (S0)0.5. exp( -((2.t1 + 2.t2)/ Tid,2D)2/3) / (1- exp( -((2.t1 + 2.t2)/ Tid,2D)2/3) )0.5 . Note here that \nR/R opt = 1-VDeer, that is why 1-VDeer is plotted on fig.6. Note also that Ropt depends o n the spin \ncoherence time T id,2D of V 2 spins probes and on the parameters t 1 and t 2 used in the \nODPELDOR exp eriment . R of course depends on the concentration of target spins C2D,T. \n \nNow, assuming a sensor operating with t 1=0.5 µs, t 2=5.75 µs and 2 t1 + 2 t2=Tid,2D =12.5µs, \nand assuming C 2D,T=1/(10nm)2 , ie suf ficiently large such that when td=5 µs, V Deer(td,C 2D,T)=0 \nie 1 - VDeer(td, dx, C2D,T)=1 (fig.6c top black curve), then one finds the simple following \nexpression for the best expected signal to noise ratio: R= (1/e) .(S0)0.5. With pex=1, pcoll=0.25 , \npdet=0.01 , a V 2 radiative recombina tion time ԎV2=6ns , and around N V2=1.5.106 V2 spins \nprobes hav ing the s ame resonant magnetic field in the sensor (see above ), and choosing a \nphotoluminescence integration time per ODPELDOR sequence T =6 µs for example, one finds \napproximately R=260, for a single \"one shot one point\" ODPELDOR experiment . The opti cal \nre-pumping time of V 2 spins is numerically evaluated to T OPump = 20 µs at 5K assuming k ISC (5K) \n=1/ (1700 ns) (see fig. 4e), but the laser pulse is assumed to last 100 µs here for safety, \nconsider ing the unmeasured value of k ISC at 5K (only known is k ISC(300K) =1/17ns at 300K31, \nsee fig. 4 d). The ODPELDOR microwave pulse s sequence after optical initialization of V 2 spins \nlast around 20 µs, such that the shot repetition time of full ODPELDOR is thus taken here to \nbe T exp=120µs. Both T tot,exp =N shot*Texp and T tot=N shot*T, increase proportionally to N shot, but R \nonly increase proportionally to (Nshot)0.5. Assuming N shot=100 per point and a 100 points \nODPELDOR spectrum as a function of f pump (1 point each 2 MHz, 200 MHz scanned), one \ncould obtain such a 200 MHz spectrum ( see fig. 6b) in 1.2 s with a signal to noise ratio \nR=2600, assuming negligible hardware and software delays for chang ing t he pumping \nmicrowave frequency (o therwise, the experimental time is determined by those delays ). \n \nIt is here also releva nt to compare standard X band direct inductively detected EPR (DD -EPR) \nsensitivity , with the one of this quantum sensor upgraded EPR method . Assuming a 2D target \nspins concentration C 2D,T=1/(10nm )2, and estimating the surface S* of target spins seen by V 2 \nspin probes and having the same resonant magnetic field to around S*=0.3 10-4 cm2 (see \nabove), one finds that around 3. 107 target spins are sensed by the V 2 spins probes in 12 ms \nper point (one point each 2 MHz, 100 shots per point), with R=2600 . As in DD -EPR15 at X \nband one can typically measure 1011 spins at 300K or 109 spins at 3K in 1 s with R DDEPR =3 \n(assuming a 1G linewidth for spins and a 1 Hz detection bandwidth) , one finds that in order \nto obtain R=2600 in 12 ms, one would need 1015 targe t spins at 300K or 1013 target s pins at \n3K with DD -EPR. Th e sensitivity gain on target spins number with this quantum sensor is thus \ncomprised between 5 and 8 orders of magnitude. Note that the probe spin s sensitivit y is SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n18 \n considerably higher than the target spins sensitivity, and it could in principle reach the single \nV2 probe spin sensitivity with long enough accumulation times . \n \n Below , I also provide some results (fig. aux. 1) of the SRIM simulation s of 22 keV As+ \nions implantation in the trilayer Zn0 (20 nm )/SiO 2(5 nm )/4H-SiC (type n <5.1015 cm-3), \nallowing, after etching of ZnO and SiO 2, to produce shallow silicon vacancies around 2 nm \nbelow the surface of 4H -SiC with an average 2D concentration of C 2D,V2= 1/(32nm)2. SRIM \nsimulations also confirms the advantage of using a trilayer and not just a Zn0/4H -SiC bilayer , \nbecause one can see on fig. aux. 2, that some Zn atoms can reach the SiO 2 layer due to the \nimplantation process and related collisions (SiO 2 is furthe r removed by etching ), but not the \nSiC substrate, thus avoiding pollution with the Zn element of the SiC substrate surface , used \nfor quantum sensing with the silicon vacancies also produced by this implantation process . \n180 200 220 240 260 280 300 320 340 360 380 400 4200,04,0x1048,0x1041,2x1051,6x105220 240 260 280 300 320 340 360 380 400 420-0,010,000,010,020,030,040,050,06As+ ions @ 22 kev \n(cm-1)\ndepth (A°) As22kevTrilVSi (per ion As+ \n@ 22 kev and per A°)\ndepth (A°) VSi\n \nfig.A ux.1: SRIM simulation of As+ ions implantation at 22 keV in this trilayer system (100 000 shots) . \n SiC-YiG X band quantum sensor . (J. Tribollet - 01/2019 ) \n19 \n \n \n \n \nfig.Aux. 2: SRIM simulation of As+ ions implantation at 22 keV in this trilayer system (here 6000 shots). \n \n Below, I also provide (fig. aux. 3) a zoom of fig. 4a used for the di scussion of \ngeometric optics in the fiber bundle based ODMR setup adapted to the SiC -YiG quantum \nsensor described here. \n \nfig.Aux. 3: Zoom of the setup for geometric optics analysis. \n \n " }, { "title": "2307.04669v2.Reversal_of_the_skyrmion_topological_deflection_across_ferrimagnetic_angular_momentum_compensation.pdf", "content": "Reversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation\nReversal of the skyrmion topological deflection across ferrimagnetic\nangular momentum compensation\nL. Berges,1R. Weil,1A. Mougin,1and J. Sampaio1\nUniversité Paris-Saclay, CNRS, Laboratoire de Physique des Solides, 91405 Orsay, France\n(*Electronic mail: joao.sampaio@universite-paris-saclay.fr)\n(Dated: 6 October 2023)\nDue to their non-trivial topology, skyrmions describe deflected trajectories, which hinders their straight propagation\nin nanotracks and can lead to their annihilation at the track edges. This deflection is caused by a gyrotropic force\nproportional to the topological charge and the angular momentum density of the host film. In this article we present\nclear evidence of the reversal of the topological deflection angle of skyrmions with the sign of angular momentum\ndensity. We measured the skyrmion trajectories across the angular momentum compensation temperature ( TAC) in\nGdCo thin films, a rare earth/transition metal ferrimagnetic alloy. The sample composition was used to engineer the\nskyrmion stability below and above the TAC. A refined comparison of their dynamical properties evidenced a reversal\nof the skyrmions deflection angle with the total angular momentum density. This reversal is a clear demonstration of\nthe possibility of tuning the skyrmion deflection angle in ferrimagnetic materials and paves the way for deflection-free\nskyrmion devices.\nThe discovery of efficient driving of chiral magnetic tex-\ntures by current-induced spin-orbit torques1–3has opened\nthe possibility of energy-efficient and high-performance spin-\ntronic devices4,5, with applications in digital6or neuromor-\nphic7–10computation, ultra-dense data-storage11,12, and sig-\nnal processing13,14. Chiral textures are stable in magnetic thin\nfilms with a significant Dzyaloshinskii-Moriya interaction\n(DMI), typically induced with an adjacent heavy-metal layer\n(e.g. Pt/Co). Additionally, the heavy-metal layer, through the\nspin Hall effect, converts an applied charge current into a spin\ncurrent that drives the magnetic textures by spin orbit torque\n(SOT). Very promising mobility of chiral magnetic domain\nwalls (DW) has been observed1,15, with nonetheless a saturat-\ning mobility at large current densities2. Another archetypal\nchiral magnetic texture is the skyrmion, a small (down to few\ntens of nm) radially symmetric whirling texture. Although\nhighly mobile16–19, their non-trivial topology induces a trans-\nverse deflection of their trajectory, a phenomenon known as\ngyrotropic deflection or skyrmion Hall effect18,20,21. This re-\nduces the velocity in the forward direction and can lead to the\nannihilation of the skyrmion at the edges of the hosting mag-\nnetic track, and is thus highly undesired.\nThe gyrotropic deflection can be mitigated in magnetic sys-\ntems with anti-parallel lattices22, such as antiferromagnets or\nferrimagnets, where the overall angular momentum density of\nthe double skyrmion can be suppressed. In particular, fer-\nrimagnetic alloys of the rare-earth/transition-metal (RETM)\nfamily, where the RE and TM moments are antiferromagnet-\nically coupled23,24, are a promising example. In a previous\nwork by our team, it was shown that skyrmions in GdCo thin\nfilms attained the high-mobility linear regime beyond pinning,\nand that their velocity and deflection followed the predictions\nof the Thiele model25. However, there is still only little ex-\nperimental evidence of the advantages of these systems26,27,\nespecially regarding the control of the gyrotropic deflection.\nIn RETMs, The balance between the moments of differ-\nent nature can be changed with alloy composition or temper-\nature which leads to two points of interest for skyrmions. At\nthe first one, the magnetic compensation temperature TMC, themagnetization of the two sub-lattices are equal, the total mag-\nnetization ( Ms=MTM−MRE) vanishes, and the size of the\nskyrmions is minimal due to the absence of dipolar fields27.\nAs RE and TM have different gyromagnetic ratios ( γREand\nγTM), the total angular momentum density ( Ls=MTM\nγTM−MRE\nγRE)\nwill vanish at a different temperature, the angular compensa-\ntion temperature TAC. Both TMCandTACdepend on composi-\ntion. The reduction and reversal of the total angular momen-\ntum, which is the root cause of magnetic precession, leads to\ninteresting dynamical properties near TAC, such as e.g. the re-\nversal of the deflection angle of chiral domain wall fingers28\nor the precessionless motion of magnetic domains walls29.\nHowever, the reversal of the skyrmion gyrotropic deflection\natTAChas not yet been demonstrated.\nIn this letter, we measure the velocity and deflection angle\nof skyrmions driven by spin-orbit torques in two Pt/GdCo/Ta\nfilms of different composition, above and below their TAC. We\nshow the dependence of the deflection with angular moment\ndensity, and in particular its reversal by changing sample com-\nposition or temperature. A quantitative analysis with a rigid\ntexture model based on the Thiele equation is used to char-\nacterize the role of the material parameters on the skyrmion\ndynamics.\nThe skyrmion dynamics were measured in two samples.\nSample 1 is composed of a film of (Si/SiOx(100))/ Ta(1)/\nPt(5)/ Gd 0.32Co0.68(5)/ Ta(3) and sample 2 of (Si/SiOx(300))/\nTa(3)/ Pt(5)/ Gd 0.3Co0.7(8)/ Ta(5)/ Pt(1) (thicknesses in nm)\nas presented in the insets in Fig. 1a. The samples were pat-\nterned into 10 µm- or 20 µm-wide tracks in order to apply\ncurrent pulses (Fig.1b). The magnetization as a function of\ntemperature was measured by SQUID magnetometry on un-\npatterned samples and is presented in Fig. 1(a). Sample 1\npresents a TMCaround 360 K whereas sample 2 presents a\nTMCaround 200 K. Therefore, at room temperature, sam-\nple 1 is RE-dominated whereas sample 2 is TM-dominated,\nwhere RE or TM domination refers to which sublattice has\nthe higher magnetic moment and therefore aligns with an ex-\nternal magnetic field. It is useful to use the effective ferromag-arXiv:2307.04669v2 [cond-mat.mtrl-sci] 5 Oct 2023Reversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation 2\nFIG. 1. a) Msversus temperature for sample 1 (top panel,\nblack points) and sample 2 (bottom panel, gray points) measured\nby SQUID (Superconducting Quantum Interference Device) magne-\ntometry, and mean-field-computed curves of Ms(solid line) and Ls\n(dashed line) for both samples. Samples stacks are presented in in-\nsets and skyrmion temperature stability regions in colored bands. b)\nTypical magnetic device studied for sample 2. c) Examples of differ-\nential MOKE images obtained in sample 2 at 350 K.\nnet model of ferrimagnets30, which assumes a signed mag-\nnetization and angular momentum density that are positive,\nby convention, when TM-dominated: Ms=|MCo|−MGd|and\nLs=|LCo|−LGd|. The exact determination of the TACis not\nstraightforward. It was therefore deduced for both samples,\nusing the mean field model described in ref.25. The calculated\nLS(T)are shown by the dashed lines in Fig. 1(a), and yield\nTAC=416 K for sample 1 and TAC=260 K for sample 2.\nThese results are consistent with the empirical law described\nin ref.31which gives TACfor GdCo between 40 to 60 K above\ntheTMC.\nThe magnetic textures are observed in each sample as\na function of temperature by magneto-optical-Kerr-effect\n(MOKE) microscopy. A typical differential MOKE image is\npresented in Fig. 1(c). Skyrmions are observed in the tem-perature ranges indicated by the color bands in Fig. 1(a). In\nthese ranges, starting from a saturated state and lowering the\napplied external magnetic field, skyrmions with a core of op-\nposing magnetization will naturally nucleate at small enough\nfield (−30 to 0 mT for an initial saturation at large negative\nmagnetic field). Skyrmions can also be nucleated by applying\nelectrical pulses25,32. A typical phase diagram (versus tem-\nperature and field) of these samples is presented in a previ-\nous work25. In the studied temperature range, sample 1 only\npresents one skyrmion stability range around 290 K, whereas\nsample 2 presents two skyrmion stability ranges, one around\n90 K and a second around 350 K. In sample 1, the skyrmion\nstability range is below TMC(and TAC), where the film LS<0,\nand so these are dubbed RE-dominated skyrmions. In sam-\nple 2, the skyrmions at 90 K are RE-dominated as well, while\nthe skyrmions at 350 K are TM-dominated (above TMCand\nTACwith therefore LS>0). Note that in the MOKE images,\nthe signal is proportional to the Co sublattice, independently\nof the temperature33. Thus, skyrmions with a core Co mo-\nment pointing along the same direction will appear with the\nsame color (black for −zwith our experimental conditions),\nwhether they are RE- or TM-dominated (Fig. 1c).\nOnce skyrmions are nucleated, electrical pulses of 3 to\n10 ns are applied and MOKE images are acquired in order\nto study the skyrmions dynamics. The skyrmion motion\nis tracked over several pulses using a partially-automated\nprocess described in ref.25, and their velocity and deflection\nare calculated considering the pulse duration and the traveled\ndistance. Typical images of skyrmions displacements are\nshown in Fig. 2, in the case of sample 2 at low temperature\nandLs<0 (a) and high temperature and Ls>0 (b). The\naverage skyrmion diameter was similar for the three studied\ncases, 0.86 ±0.28 µm. An example of the observed skyrmion\ndynamics in sample 1 is presented in Fig. 2(c) with a super-\nposition of successive MOKE images where the skyrmion\ncolor refers to the MOKE image number.\nThe skyrmion deflection ( θsk) and velocity ( v) versus\napplied current density ( j) are presented in Fig. 3(a,b) for\nthe three cases: RE-dominated skyrmions in sample 1, and\nRE- and TM-dominated skyrmions in sample 2. Videos of\nsuccessive displacements in both samples are shown in S.I. In\nthe three cases, the velocity shows a clear depinning transition\nabove a current threshold (different for each case), and then\nfollows a linear regime. The mobility in the linear regime\n(i.e.∆v/∆j) is much higher in sample 2 than in sample 1. In\nsample 2, the mobility of TM-dominated skyrmions is slightly\nhigher than RE-dominated skyrmions. These differences in\nmobility will be discussed later. The linear regime extends up\nto 190 m/s in sample 1 and to 450 m/s in sample 2. At highest\nj, skyrmions are nucleated by the pulse, which hinders the\ntracking analysis and thus limits the maximum jthat can\nbe examined. In the linear regime, the deflection angle θsk\nis approximately constant with the current density, and its\nabsolute value is about 40◦for the three cases. The deflection\nangle is clearly reversed between the TM- and RE-dominated\nskyrmions: it is positive for TM-dominated skyrmions (in\nsample 2) and negative for RE-dominated skyrmions (in bothReversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation 3\nFIG. 2. a)-b) Example of three successive MOKE images separated by 10-ns (6 ns) pulses showing the displacement of skyrmions in sample\n2 with at a) at T = 110 K and j=120 GA/m2and b) T = 350 K and 150 GA/m2(field around 0 mT.) The colored circles identify the same\nskyrmions in the three images. As the images in a) were obtained using a cryostat and are of lower resolution, a different temperature from\nthe one used in the dynamical studies (90 K) was used to render the skyrmions larger and more visible. c) Superposition of four consecutive\nMOKE images, in the case of sample 1 with Ls<0, showing the propagation of three skyrmions for 2 and 4 images. The different forces\ndefined in the Thiele equation acting on the skyrmions are sketched around the black dot.\nFIG. 3. a) Averaged skyrmion deflection θsk, and b) velocity, mea-\nsured in sample 1 at 290 K and 2 at 90 K and 350 K. The error bars\nrepresent the standard deviation of the measurements. The dashed\nlines are obtained with the Thiele based model (with fitted θSHE).\nThe color bands represent the expected error due to the experimental\nerror of θSHE, as described in the main text.\nsamples). The deflection also reverses with core polarity,\ni.e. with the Co moment pointing along +z(which appear\nas white skyrmions in the MOKE images; see Supplemental\nMaterials). The θskin the pining regime is measured to belarger than in the flow regime in sample 1, whereas it is lower\nin sample 2. This is perhaps a bias induced by the different\nnucleation protocol used in these measurements. For sample\n1, skyrmions were only nucleated by current pulses, mostly\nnear one of the edges due to the Oersted field25, whereas\nfor sample 2 they were first nucleated homogeneously by\nmagnetic field. As skyrmions can be annihilated at the\nedges, only the skyrmions that deviate towards the center are\naccounted for, which biases the measurement of the mean θsk.\nThe skyrmion dynamics in the linear regime can be quanti-\ntatively analyzed using a rigid-texture formalism based on the\nThiele equation34. It expresses the equilibrium of all forces\napplied on the magnetic texture that reads in our case as:\nFG+FSOT+αDv=0, where FSOTis the SOT force, FGthe\ngyrotropic force and αD is, in general, a tensor describing\nthe dissipation. This formalism can be applied to skyrmions\nin double-lattice systems as presented in refs.25,35. These\nforces are depicted in Fig. 2 c), on a black dot representing\na skyrmion in the case of Ls<0. The norm of the skyrmion\nvelocity |v|and its deflection θskcan be deduced to be:\n|v|=v0p\n1+ρ2(1)\nθsk=arctan (ρ) (2)\nIn the limit of skyrmions larger than the domain wall width\nparameter ∆, the parameters v0andρare:\nv0≈ −π∆\n2Lα¯h jθSHE\n2et(3)\nρ≈∆\n2πRLS\nLαn (4)Reversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation 4\nTTAC\nMs − + +\nLs − − +\ncore magnetic moment +z⊙ − z⊗ − z⊗\ncore cobalt moment −z⊗ − z⊗ − z⊗\npCo −1 −1 −1\nFG·y − − +\nFSOT +x +x +x\nTABLE I. Signs versus temperature of the material parameters,\nof the skyrmion core configuration parameters and of the expected\nforces acting on a skyrmion with negative polarity (black in the\nMOKE images) driven by a positive ( +x) current.\nParameter Sample 1 (290 K) Sample 2 (350 K)\nγ/2π[GHz/T]a18.3 40.8\nαa0.15 0.175\n|MS|[kA/m]b78 125\n|LS|[kg/(ms)]b6.8×10−74.9×10−7\n|Lα|[kg/(ms)]b1.02×10−78.5×10−8\nµ0Hk[mT]c200 60\nKu[kJ/m3]c11.5 13.6\nD[ mJ/m2]a−0.22 −0.14\nA[pJ/m]a4.6 (4.6)d\nθSHEe0.04 0.09\n2R[µm]f0.85±0.28 0 .86±0.28\nTABLE II. Measured parameters used in the model.aγ,α, the ex-\nchange stiffness Aand the DMI strength D(not used in the model)\nwere determined with Brillouin light scattering (BLS) at 290 K (sam-\nple 1) or 350 K (sample 2).bMs(T)(Fig. 1a) was used to determine\nLS=MS/γandLα=LSα.cHkwas measured from hysteresis cy-\ncles and BLS, from which Kuwas deduced.dAwas measured only\non sample 1 and assumed to be the same in sample 2.eθSHEwas\ndetermined by transport measurements using the double-harmonic\ntechnique25,36(see Supplementary Materials).fThe shown variation\nis the standard deviation of the observed radius and not the error of\nthe average value.\nwhere ¯his the Planck constant, ethe fundamental charge,\ntthe magnetic film thickness, θSHE is the effective SHE\nangle in the Pt layer, Lα=LSαthe energy dissipation rate,\nn=pCo4π=±4πthe topological charge of the skyrmion, R\nits radius, and pCo=±1 is the orientation along zof the core\nCo moment. Because Lαis always positive, the sign of the\ndeflection is given by the sign of the product of Ls(positive\nforT>TAC) and pCo. This sign is presented in Table I as\na function of temperature for pCo=−1, which is the case\nshown here (black skyrmions).\nThe parameters needed for the model were measured on\nboth samples (see Table II). Hk(T)was obtained by analyz-\ning hysteresis loops at various temperatures, which yielded a\nvalue for Ku(Ku=µ0HkMs/2−µ0M2\ns/2) with negligible ther-\nmal variation. The domain wall width parameter was calcu-\nlated using ∆=p\nA/Keff, where Keff=µ0HkMs/2 is the effec-\ntive anisotropy. For Sample 2 at 90 K ( MS=135 kA/m), the\nthermal variation of KuandAwas neglected (as it is smallerthan the precision of the other parameters) and the values at\n350 K were used; LS=12.1×10−7kg/(ms) and Lα=10.7\n×10−8kg/(ms) were deduced using a mean-field model as de-\nscribed in ref.25, assuming constant sub-lattice Gilbert damp-\ning parameters ( Lα=αCo|LCo\ns(T)|+αGd|LGd\ns(T)|). The\nskyrmion diameter was taken from the average observed di-\nameter, which is very similar for the three studied cases37.\nThese measured parameters allow to constrain the model\nand obtain curves for the velocity and deflection angle (dashed\nlines in Fig. 3). A constant deflection is predicted, and its\nvalue is obtained with no fitting parameters. The velocity is\npredicted to be linear with j, and its slope is obtained with\na single fitting parameter, θSHE. The fitted values ( θSHE =\n0.03 for sample 1 and 0.09 for sample 2) are consistent with\nthe precision of the measured θSHE(see Table II and Suppl.\nMat.). The model prediction range, calculated with the esti-\nmated error θSHE, is shown in the figure as a color band38.\nAbove the depinning threshold, where the model is expected\nto be valid, it both reproduces the qualitative behavior (con-\nstant deflection and linear velocity) and agrees quantitatively\nwith the experimental data, within the estimated error margin.\nIn particular, the sign of the deflection angle observed in the\nexperiments agrees with Eq (3b) taking into account the LSof\nthe film ( LS<0 for RE-dominated skyrmions and LS>0 for\nTM-dominated skyrmions).\nThe skyrmion mobility, given by the slope of the velocity\nversus j(Fig. 3(b)), is much higher in sample 2 than in sample\n1 (1.80 at 350 K vs 0.6 m ·s−1/GA·m−2, respectively). This\ndifference in mobility cannot be ascribed to a difference in\nskyrmion diameter (see eq. 1), as the two conditions present\nvery similar average sizes (Table II). This large difference has\nmultiple origins. First, the Lαof sample 2 is lower by 20%.\nThe second major cause is the difference of the film stacks, in\nparticular the thickness of the Ta capping layer. The measured\nθSHEis more than twice higher in sample 2 than in sample 1\n(Table II). This can be expected to be due a better passivation\nof the Ta layer in sample 2 which can therefore contribute\nmore to the SOT than the thinner (3 nm) Ta cap of the sample\n1 which is probably fully oxidized.\nFinally, comparing the skyrmion velocity curves for the\ntwo conditions in sample 2 (at 90 and 350 K), it can be seen\nthat both the depinning current and the mobility in the linear\nregime are significantly different. The depinning current is\nhigher at 90 K, which can be attributed by the thermal nature\nof the depinning process39. The difference in mobility is not\ndue to a difference in skyrmion diameter (which again is very\nsimilar in all three studied conditions). It can be expected that\nseveral magnetic parameters vary between 90 and 350 K, but\nthe experimental mobility can be understood by considering\nonly the variation of Lα(Lα(90 K )\nLα(350 K )≈1.25). This result and the\nThiele model suggest that Lαis a more pertinent parameter\nthan αto characterize the role of dissipation in the skyrmion\nmobility. Interestingly, Lαcan be more easily optimized than\nαto increase mobility, by increasing the sample temperature\n(as was the case here) or by decreasing the material’s Curie\ntemperature (all other parameters remaining equal). A recent\nwork39on skyrmions measured at relatively high temperature\nalso seems to point toward such an effect which seems to beReversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation 5\nan interesting path to increase skyrmion mobility.\nIn conclusion, we observed the propagation of skyrmions\nin the flow regime, i.e., beyond the effects of pinning in two\nGdCo samples, below and above the angular compensation\ntemperature. The observed mobilities were very large, with\na velocity up to 450 m/s. The skyrmion dynamics was stud-\nied in three cases, two in RE-dominated films and one in a\nTM-dominated film. The deflection angle was constant with\ndriving current and its sign was opposite between RE- and\nTM-dominated cases, both when comparing two samples of\ndifferent composition and when comparing two temperatures\n(above and below TAC) in the same sample. This confirms the\nmodulation of deflection angle θskwith LS.\nThese experiments demonstrate the effects of the angular\nmomentum density LSof the host material on the deflection of\nskyrmions. They show that θskcan be reversed in GdCo ferri-\nmagnetic thin films across their angular compensations, either\nby changing the alloy stoichiometry or simply its temperature.\nIn particular, the reversal of sign of θskacross compensation\nstrongly supports that θskshould be zero at angular moment\ncompensation. The engineering of magnetic parameters that\nwas done to produce the two presented skyrmion-hosting sam-\nples could be repeated rather straightforwardly to engineer a\nfilm with stable skyrmions at TACwith no deflection.\nACKNOWLEDGMENTS\nThe authors thank Stanislas Rohart for fruitful discussions,\nand André Thiaville for the study of the sample properties by\nBLS. This work was supported by a public grant overseen by\nthe French National Research Agency (ANR) as part of the\n“Investissements d’Avenir” program (Labex NanoSaclay, ref-\nerence: ANR-10-LABX-0035, project SPICY). Magnetome-\ntry and Anomalous Hall effect measurements were performed\nat the LPS Physical Measurements Platform.\nSUPPLEMENTARY MATERIAL\nSee supplementary material for videos of successive\nMOKE images showing the skyrmion motion for the three\ntemperature regions discussed in the text. Motion of\nskyrmions of opposite polarity (i.e., pCo= +1; white in the\nMOKE images) is also shown for sample 1. A note on the\nanalysis of the images and the selection of relevant textures\nis also included. Experimental data of θSHEare also included\nfor both samples at several temperatures.\nDATA AVAILABILITY STATEMENT\nThe data that support the findings of this study are available\nfrom the corresponding author upon reasonable request.1T. A. Moore, I. M. Miron, G. Gaudin, G. Serret, S. Auffret, B. Rodmacq,\nA. Schuhl, S. Pizzini, J. V ogel, and M. Bonfim, “High domain wall veloc-\nities induced by current in ultrathin Pt/Co/AlOx wires with perpendicular\nmagnetic anisotropy,” Applied Physics Letters 93, 262504 (2008).\n2A. Thiaville, S. Rohart, É. Jué, V . Cros, and A. Fert, “Dynamics of\nDzyaloshinskii domain walls in ultrathin magnetic films,” EPL (Euro-\nphysics Letters) 100, 57002 (2012).\n3A. Manchon, J. Železný, I. M. Miron, T. Jungwirth, J. Sinova, A. Thiaville,\nK. Garello, and P. Gambardella, “Current-induced spin-orbit torques in\nferromagnetic and antiferromagnetic systems,” Rev. Mod. Phys. 91, 035004\n(2019).\n4A. Fert, N. Reyren, and V . Cros, “Magnetic skyrmions: advances in physics\nand potential applications,” Nature Reviews Materials 2, 17031 (2017).\n5J. Sampaio, V . Cros, S. Rohart, A. Thiaville, and A. Fert, “Nucleation, sta-\nbility and current-induced motion of isolated magnetic skyrmions in nanos-\ntructures,” Nature Nanotechnology 8, 839–844 (2013).\n6S. Zhang, A. A. Baker, S. Komineas, and T. Hesjedal, “Topological com-\nputation based on direct magnetic logic communication,” Scientific Reports\n5, 15773 (2015).\n7Y . Huang, W. Kang, X. Zhang, Y . Zhou, and W. Zhao, “Magnetic\nskyrmion-based synaptic devices,” Nanotechnology 28, 08LT02 (2017).\n8J. Zázvorka, F. Jakobs, D. Heinze, N. Keil, S. Kromin, S. Jaiswal, K. Litz-\nius, G. Jakob, P. Virnau, D. Pinna, K. Everschor-Sitte, L. Rózsa, A. Donges,\nU. Nowak, and M. Kläui, “Thermal skyrmion diffusion used in a reshuffler\ndevice,” Nature Nanotechnology 14, 658–661 (2019).\n9S. Li, W. Kang, Y . Huang, X. Zhang, Y . Zhou, and W. Zhao, “Magnetic\nskyrmion-based artificial neuron device,” Nanotechnology 28, 31LT01\n(2017).\n10K. M. Song, J.-S. Jeong, B. Pan, X. Zhang, J. Xia, S. Cha, T.-E. Park,\nK. Kim, S. Finizio, J. Raabe, J. Chang, Y . Zhou, W. Zhao, W. Kang, H. Ju,\nand S. Woo, “Skyrmion-based artificial synapses for neuromorphic com-\nputing,” Nature Electronics 3, 148–155 (2020).\n11A. Fert, V . Cros, and J. Sampaio, “Skyrmions on the track,” Nature Nan-\notechnology 8, 152–156 (2013).\n12A. Brataas, A. D. Kent, and H. Ohno, “Current-induced torques in magnetic\nmaterials,” Nature Materials 11, 372–381 (2012).\n13M. Carpentieri, R. Tomasello, R. Zivieri, and G. Finocchio, “Topological,\nnon-topological and instanton droplets driven by spin-transfer torque in ma-\nterials with perpendicular magnetic anisotropy and Dzyaloshinskii-Moriya\nInteraction,” Scientific Reports 5, 1–8 (2015).\n14G. Finocchio, M. Ricci, R. Tomasello, A. Giordano, M. Lanuzza, V . Puli-\nafito, P. Burrascano, B. Azzerboni, and M. Carpentieri, “Skyrmion based\nmicrowave detectors and harvesting,” Applied Physics Letters 107, 3–8\n(2015).\n15K.-J. Kim, S. K. Kim, Y . Hirata, S.-H. Oh, T. Tono, D.-H. Kim, T. Okuno,\nW. S. Ham, S. Kim, G. Go, Y . Tserkovnyak, A. Tsukamoto, T. Moriyama,\nK.-J. Lee, and T. Ono, “Fast domain wall motion in the vicinity of the an-\ngular momentum compensation temperature of ferrimagnets,” Nature Ma-\nterials 16, 1187–1192 (2017).\n16O. Boulle, J. V ogel, H. Yang, S. Pizzini, D. de Souza Chaves, A. Locatelli,\nT. O. Mente¸ s, A. Sala, L. D. Buda-Prejbeanu, O. Klein, M. Belmeguenai,\nY . Roussigné, A. Stashkevich, S. Mourad Chérif, L. Aballe, M. Foerster,\nM. Chshiev, S. Auffret, I. M. Miron, G. Gaudin, S. M. Chérif, L. Aballe,\nM. Foerster, M. Chshiev, S. Auffret, I. M. Miron, and G. Gaudin, “Room-\ntemperature chiral magnetic skyrmions in ultrathin magnetic nanostruc-\ntures,” Nature Nanotechnology 11, 449–454 (2016).\n17A. Hrabec, J. Sampaio, M. Belmeguenai, I. Gross, R. Weil, S. M. Chérif,\nA. Stashkevich, V . Jacques, A. Thiaville, and S. Rohart, “Current-induced\nskyrmion generation and dynamics in symmetric bilayers,” Nature Com-\nmunications 8, 15765 (2017).\n18W. Jiang, X. Zhang, G. Yu, W. Zhang, X. Wang, M. Benjamin Jungfleisch,\nJ. E. Pearson, X. Cheng, O. Heinonen, K. L. Wang, Y . Zhou, A. Hoffmann,\nand S. G. E. te Velthuis, “Direct observation of the skyrmion Hall effect,”\nNature Physics 13, 162–169 (2017).\n19C. Reichhardt, C. J. O. Reichhardt, and M. V . Miloševi ´c, “Statics and dy-\nnamics of skyrmions interacting with disorder and nanostructures,” Rev.\nMod. Phys. 94, 035005 (2022).\n20J. Zang, M. Mostovoy, J. H. Han, and N. Nagaosa, “Dynamics of skyrmion\ncrystals in metallic thin films,” Phys. Rev. Lett. 107, 136804 (2011).\n21G. Chen, “Skyrmion hall effect,” Nature Physics 13, 112–113 (2017).Reversal of the skyrmion topological deflection across ferrimagnetic angular momentum compensation 6\n22T. Dohi, S. DuttaGupta, S. Fukami, and H. Ohno, “Formation and current-\ninduced motion of synthetic antiferromagnetic skyrmion bubbles,” Nature\nCommunications 10, 5153 (2019).\n23P. Hansen, C. Clausen, G. Much, M. Rosenkranz, and K. Witter, “Magnetic\nand magneto-optical properties of rare-earth transition-metal alloys con-\ntaining Gd, Tb, Fe, Co,” Journal of Applied Physics 66, 756–767 (1989).\n24G. Sala and P. Gambardella, “Ferrimagnetic Dynamics Induced by Spin-\nOrbit Torques,” Advanced Materials Interfaces 2201622 , 2201622 (2022).\n25L. Berges, E. Haltz, S. Panigrahy, S. Mallick, R. Weil, S. Rohart, A. Mou-\ngin, and J. Sampaio, “Size-dependent mobility of skyrmions beyond pin-\nning in ferrimagnetic GdCo thin films,” Physical Review B 106, 144408\n(2022).\n26S. Woo, K. M. Song, X. Zhang, Y . Zhou, M. Ezawa, X. Liu, S. Finizio,\nJ. Raabe, N. J. Lee, S.-I. Kim, S.-Y . Park, Y . Kim, J.-Y . Kim, D. Lee, O. Lee,\nJ. W. Choi, B.-C. Min, H. C. Koo, and J. Chang, “Current-driven dynamics\nand inhibition of the skyrmion Hall effect of ferrimagnetic skyrmions in\nGdFeCo films,” Nature Communications 9, 959 (2018).\n27L. Caretta, M. Mann, F. Büttner, K. Ueda, B. Pfau, C. M. Günther, P. Hess-\ning, A. Churikova, C. Klose, M. Schneider, D. Engel, C. Marcus, D. Bono,\nK. Bagschik, S. Eisebitt, and G. S. D. Beach, “Fast current-driven do-\nmain walls and small skyrmions in a compensated ferrimagnet,” Nature\nNanotechnology 13, 1154–1160 (2018).\n28Y . Hirata, D.-H. Kim, S. K. Kim, D.-K. Lee, S.-H. Oh, D.-Y . Kim,\nT. Nishimura, T. Okuno, Y . Futakawa, H. Yoshikawa, A. Tsukamoto,\nY . Tserkovnyak, Y . Shiota, T. Moriyama, S.-B. Choe, K.-J. Lee, and\nT. Ono, “Vanishing skyrmion Hall effect at the angular momentum compen-\nsation temperature of a ferrimagnet,” Nature Nanotechnology 14, 232–236\n(2019).\n29E. Haltz, J. Sampaio, S. Krishnia, L. Berges, R. Weil, and A. Mou-\ngin, “Measurement of the tilt of a moving domain wall shows precession-\nfree dynamics in compensated ferrimagnets,” Scientific Reports 10, 16292\n(2020).\n30R. K. Wangsness, “Sublattice Effects in Magnetic Resonance,” Physical Re-\nview 91, 1085–1091 (1953).31Y . Hirata, D.-H. Kim, T. Okuno, T. Nishimura, D.-Y . Kim, Y . Futakawa,\nH. Yoshikawa, A. Tsukamoto, K.-J. Kim, S.-B. Choe, and T. Ono, “Cor-\nrelation between compensation temperatures of magnetization and angu-\nlar momentum in GdFeCo ferrimagnets,” Physical Review B 97, 220403\n(2018).\n32Y . Quessab, J.-W. Xu, E. Cogulu, S. Finizio, J. Raabe, and A. D. Kent,\n“Zero-field nucleation and fast motion of skyrmions induced by nanosecond\ncurrent pulses in a ferrimagnetic thin film,” Nano Letters 22, 6091–6097\n(2022).\n33L. Berges, “Magnetic skyrmions in gdco ferrimagnetic thin-films,” (2022),\nphD thesis defended at université Paris-Saclay.\n34A. A. Thiele, “Applications of the gyrocoupling vector and dissipation\ndyadic in the dynamics of magnetic domains,” Journal of Applied Physics\n45, 377–393 (1974).\n35S. Panigrahy, S. Mallick, J. Sampaio, and S. Rohart, “Skyrmion inertia in\nsynthetic antiferromagnets,” Physical Review B 106, 144405 (2022).\n36M. Hayashi, J. Kim, M. Yamanouchi, and H. Ohno, “Quantitative char-\nacterization of the spin-orbit torque using harmonic Hall voltage measure-\nments,” Physical Review B 89, 144425 (2014).\n37The skyrmions show a large dispersion of diameter (Table II), which we\nestimate to lead to a ±14% dispersion of velocity and to a ±10º dispersion\nof the deflection angle. However, as the values of velocity and deflection\nwere averaged over many skyrmions and many displacements, the error\ndue to the size dispersion is drastically reduced and is neglected.\n38We estimate that the main sources of error of ρandvis the spin Hall effi-\nciency ( θSHE). which was estimated to be 25% (sample 1) and 12% (sample\n2); see supplemental materials.\n39K. Litzius, J. Leliaert, P. Bassirian, D. Rodrigues, S. Kromin, I. Lemesh,\nJ. Zazvorka, K.-J. Lee, J. Mulkers, N. Kerber, D. Heinze, N. Keil, R. M.\nReeve, M. Weigand, B. Van Waeyenberge, G. Schütz, K. Everschor-Sitte,\nG. S. D. Beach, and M. Kläui, “The role of temperature and drive current\nin skyrmion dynamics,” Nature Electronics 3, 30–36 (2020)." }, { "title": "1506.02532v2.Geometric__electronic_and_magnetic_structure_of_Fe___x__O___y_______clusters.pdf", "content": "arXiv:1506.02532v2 [physics.atm-clus] 20 Nov 2015Geometric, electronic, and magnetic structure of FexO+\nyclusters\nR. Logemann, G.A. de Wijs, M.I. Katsnelson, and A. Kirilyuk\nRadboud University, Institute for Molecules and Materials , NL-6525 AJ Nijmegen, The Netherlands\n(Dated: August 5, 2021)\nCorrelation between geometry, electronic structure and ma gnetism of solids is both intriguing and\nelusive. This is particularly strongly manifested in small clusters, where a vast number of unusual\nstructures appear. Here, we employ density functional theo ry in combination with a genetic search\nalgorithm, GGA+ Uand a hybrid functional to determine the structure of gas pha se FexO+/0\nyclus-\nters. For Fe xO+\nycation clusters we also calculate the corresponding vibrat ion spectra and compare\nthem with experiments. We successfully identify Fe 3O+\n4, Fe4O+\n5, Fe4O+\n6, Fe5O+\n7and propose struc-\ntures for Fe 6O+\n8. Within the triangular geometric structure of Fe 3O+\n4a non-collinear, ferrimagnetic\nand ferromagnetic state are comparable in energy. Fe 4O+\n5and Fe 4O+\n6are ferrimagnetic with a resid-\nual magnetic moment of 1 µBdue to ionization. Fe 5O+\n7is ferrimagnetic due to the odd number of\nFe atoms. We compare the electronic structure with bulk magn etite and find Fe 4O+\n5, Fe4O+\n6, Fe6O+\n8\nto be mixed valence clusters. In contrast, in Fe 3O+\n4and Fe 5O+\n7, all Fe are found to be trivalent.\nPACS numbers: 36.40.Cg, 36.40.Mr, 61.46.Bc, 73.22.-f\nIn nano technology there is an ever increasing demand\nfor increasing the density of electronic and magnetic de-\nvices. This continuous downscaling trend drives the in-\nteresttoelectronicandmagneticstructuresatthe atomic\nscale. In essence, two things are required: first, novel\nmaterials and building blocks with exotic physical prop-\nerties. Second, a fundamental knowledge of the physical\nmechanism of magnetism at the sub-nanometer scale.\nAtomicclusters,havinghighlynon-monotonousbehav-\nior as a function of size, are a promising model system to\nstudy the fundamentals of magnetism at the nanoscale\nand below. Such clusters consist of only tens of atoms.\nQuantum mechanics starts to play an essential role at\nthis small scale, adding extra degrees of freedom. Since\nthese clusters are studied in high vacuum, they are com-\npletely isolated from their environment.\nTo use these clusters as a model system, as a starting\npoint, a detailed understanding of the relation between\ntheir geometry and electronic structure is required.\nEvenin thebulk, ironoxidehasawidevarietyofchem-\nical compositions and phases with many interesting phe-\nnomena, such as the Verwey transition in magnetite.1,2\nExperimentsperformedonsmallgasphaseFe xOyclus-\nters beyond the two-atom case are scarce. The structure\nof one and two Fe atoms with oxygen has been studied\nin an argon matrix using infrared spectra.3,4The corre-\nsponding vibration frequencies have been identified using\ndensity functional theory (DFT).\nIron-oxide nanoparticles have been investigated for\ntheirpotentialuseascatalystinchemicalreactions.5Fur-\nthermore, since the iron-oxygen interaction has a funda-\nmental role in many chemical and biological processes,\nthere have been quite some studies, both experimen-\ntal and theoretical, of the chemical properties of Fe xOy\nclusters.6–12\nThe possible coexistence of two structural isomers for\nstoichiometric iron-oxide clusters in the size range n≥5\nwas experimentally measured using isomer separation by\nion mobility mass spectroscopy for FenOnand FenOn+1(n= 2-9).13Furthermore, the formation of Fe xOyclus-\nters has been studied in the size range ( x= 1-52).14\nThe number of theoretical studies is, however, man-\nifold. The magic cluster Fe 13O8was extensively stud-\nied and identified as a cluster with C1but close to D4h\npoint group symmetry.15–19However, also the geometry\nand electronic structure of other cluster sizes have been\nstudied theoretically.15,20–25The prediction of geometric\nstructures requires a systematic search of the potential\nenergy surface to find the global minimum.\nThe majority of theoretical studies were performed\nusing DFT.4,6,9,10,13–17,20–24,26The number of works in\nwhichFe mOnclusterswerestudied with methods beyond\nDFT is very limited and restricted to very small cluster\nsizes. For FeO+its reactivity towards H2was studied\non a wave-function-based CASPT2D level.12For Fe2O2\nthemolecularandelectronicstructurewerecalculatedus-\ning both DFT and wave-function-based CCSD(T) meth-\nods and a7B2uground state was found.25Furthermore,\nRef. 25 reports that B3LYP functional and CCSD(T)\ncalculations give the same energy ordering of different\nstates, although the energy differences are overestimated\nby the B3LYP approach.\nRecently, the structural evolution of (Fe 2O3)n\nnanoparticles was systematically investigated from the\nFe2O3cluster towards nano particles with n= 1328.9,26\nIn the size range of n= 1-10, an interatomic potential\nwas developed and combined with a genetic algorithm in\nsearchofthe lowest-energyisomer. The isomerslowest in\nenergy were further optimized using DFT and the hybrid\nfunctional B3LYP. This way, a systematic prediction of\nthe cluster structure was done for neutral (Fe 2O3)nclus-\nters.\nBecause of its high computational burden, in DFT the\ngeometric structure is often only relaxed into its nearest\nlocal minimum on the potential energy surface (PES).\nThere is no guarantee that this local minimum corre-\nsponds to the global minimum. Almost all previous2\nworksonlyconsidereitherrandomstructuresormanually\nconstructed geometries. However, for increasing cluster\nsize these methods become less successful in finding the\nlowest-energy isomer. Genetic algorithms, in which sta-\nble geometries are used to create new structures, proved\nto be efficient in finding the global energy minimum.27\nThis method has been successfully used for transition-\nmetal oxide clusters.28,29\nIdentification of the geometric cluster structure is a\ndelicate and computationally demanding task. There-\nfore,comparisonwithanexperimentalmethodtoconfirm\nthe theoreticalfindingsis essential. In this work, wecom-\nbine previously reported experimental vibration spec-\ntra30with first-principles calculations and a genetic al-\ngorithm to determine the geometric structure of cationic\nFexO+\nyclusters. Of the nine cluster sizes reported in\nRef. 30, only the geometric structure of Fe 4O+\n6was iden-\ntified. In this work, we will also identify the geomet-\nric, electronic, and magnetic structure of Fe 3O+\n4, Fe4O+\n5,\nFe5O+\n7and propose structures for Fe 6O+\n8.\nI. COMPUTATIONAL DETAILS\nWe employ a genetic algorithm (GA) as is described in\nRef. 27 in combination with DFT to optimize the cluster\nstructures. For this we use the Vienna ab-initio simu-\nlation package ( vasp)31using the projector augmented\nwave (PAW) method.32,33Since the geometry optimiza-\ntion is the most computationally expensive part of the\ngenetic algorithm, we use the PBE+ Umethod34with\nlimited accuracy for the genetic algorithm. For all ob-\ntained isomers low in energy, we reoptimized the geo-\nmetric structure using the hybrid B3LYP functional with\nhigher accuracy and consider different magnetic config-\nurations. We then calculate the vibration spectra and\ncompare them with experimental results.\nWithin the DFT framework, functionals based on the\nlocal density approximation (LDA) or general gradient\napproximation (GGA) fail to describe strongly interact-\ning systems such as transition-metal oxides.35,36Due to\nthe overestimation of the electron self-interaction, they\npredict metallic behavior instead of the (correct) wide-\nband-gap insulator. In an attempt to correct for this\nself-interaction, one can, for example, employ a hybrid\nfunctional, where a typical amount of 20% of Hartree-\nFockenergyisincorporatedintotheexchange-correlation\nfunctional. Especially for the B3LYP functional it has\nbeen shown that this results in good agreement be-\ntween the geometric structure and vibrational spectra\nfor clusters.28,30,37However, hybrid functionals are quite\ncomputationally expensive compared to LDA and GGA\nfunctionals. Therefore, in the genetic algorithm we em-\nploy the GGA+ Umethod to take into account that FeO\nclusters are strongly interacting systems. We use the\nrotational invariant implementation introduced by Du-\ndarev and a plane wave cutoff energy of 300 eV for these\ncalculations.38The differences between GGA and GGA+ Ufor iron-\noxide cluster calculations have been analyzed in Ref. 15.\nThisstudystressestheimportancetogobeyondGGAfor\ntransition-metal oxide clusters calculations. Aside from\nthe well-known difference for the electronic and magnetic\nstructure, it even finds a different lowest energy isomer\nthan GGA for Fe 32O33. In our genetic algorithm cal-\nculations we use an Ueff=U−Jof 3 eV for the Fe\natoms, based on a comparison between B3LYP calcula-\ntions and PBE+ Ucalculations for the smallest cluster,\nFe3O4(see Sec. IIB). For this comparison we also cal-\nculated the mean absolute difference (∆) between the\noccupied Kohn-Sham energies ( Ei) using B3LYP and\nPBE+U:\n∆ =n/summationdisplay\ni=1|EPBE+U\ni−EB3LYP\ni|\nn, (1)\nwherenis the number of occupied Kohn-Sham levels.\nNote that, the binding distances are only weakly depen-\ndent on the used Ueffand our value of 3 eV is close to val-\nues used in other works (e.g., 5 eV15, 3.6 eV20, 3.6 eV39).\nWe used the genetic algorithm as described in detail\nin Ref. 27. New geometries are formed by the Deaven-\nHo cut and splice crossover operation. To determine the\nfitness we used an exponential function. A generation\ntypically consists of 20 clusters. It has been shown that\nthe geometry of Fe xOyclusters only weakly depends on\nthe magnetic degree of freedom.26Therefore, we restrict\nourselves to the ferromagnetic case in our genetic algo-\nrithm.\nFor all obtained isomers low in energy, we reopti-\nmized the geometric structure using the hybrid B3LYP\nfunctional40,53and consider all possible collinear ori-\nentations of the Fe magnetic moments by constraining\nthe difference in majority and minority electrons. All\nforces were minimized below 10−3eV/˚A. Standard rec-\nommended PAWs with an energy cutoff of 400.0 eV are\nused. The clusters are placed in a periodic box of a\nsize between 11 and 17 ˚A, which we checked to be suf-\nficiently large to eliminate inter cluster interactions for\neach cluster size. For the cluster calculations, a single\nk-point (Γ) is used. Since we also consider cationic clus-\nters, a positive uniform background charge is added and\nwe correct the leading errors in the potential.41,42All\nsimulations were performed without any symmetry con-\nstraints. The reported symmetry groups are determined\nafterwardswithin 0.03 ˚A. Forthe density ofstates(DOS)\ncalculations we used a Gaussian smearing of 0.1 eV for\nvisual clarity.\nTo obtain the vibration spectra, the Hessian matrix\nof an optimized geometry is calculated by considering\nfinite ionic displacements of 0.015 ˚A for all Cartesian co-\nordinates of each atom. The vibration frequencies are\nobtained by diagonalization of the Hessian matrix. The\nabsorption intensity Aiis calculated using43,44\nAi= 974.86gi/parenleftbigg∂µ\n∂Qi/parenrightbigg\n, (2)3\nwheregiis the degeneracy of the vibration mode, Qi\nthe mass weighted vibrational mode, µthe electric dipole\nmoment, and974.86anempiricalfactor. Amethodbased\non four displacements for each ion was also tested but\nyielded the same frequencies and absorption intensities.\nZero-point vibrational energies (ZPVE) were calculated\nfor the isomers lowest in energy of which the vibration\nspectra are also shown.\nFor a quantitative comparison between experimen-\ntal and calculated vibrational spectra, we calculate the\nPendry’s reliability factor.45The Pendry’s reliability fac-\ntor is a well-established method in low-energy electron\ndiffraction (LEED) to quantify the agreement in contin-\nuous spectra and has also been applied to vibrational\nspectroscopy.46\nThe experimental used infrared multiphoton dissoci-\nation method (IR-MPD) does not only depend on the\nabsorption cross section of a vibrational mode, but also\non the dissociation cross section. Therefore, we use the\nPendry’s reliability factor to quantify the comparison of\nvibration spectrasince it is mainly sensitive to peak posi-\ntions opposed to a comparison of squared intensity. This\npeak sensitivity is achieved by comparing the renormal-\nized logarithmic derivative of the intensity I(ω):\nY(ω) =L−1(ω)\nL−2(ω)+W2, (3)\nwhereL(ω) =I′(ω)/I(ω) andWis the typicalFWHM of\nthe peaks in the spectra. The Pendry’s reliability factor\nis defined as:\nRP=/integraldisplay/bracketleftbig\nYth(ω)−Yexpt(ω)/bracketrightbig2\nY2\nth(ω)+Y2\nexpt(ω)dω, (4)\nwhere we integrate over the experimental range of fre-\nquencies. RPvalues range from 0 to 2, where 0 means\nperfect agreement, 1 uncorrelated spectra, and 2 per-\nfect anticorrelation. In practice, RPvalues of 0.3 are\nconsidered acceptable agreement within LEED. Y(ω) is\nstrongly dependent on experimental noise and values\nclose to zero, hence, we calculate Yexpt(ω) by fitting the\nexperimental spectrum with multiple Lorentzian peaks\nand extract the corresponding W. The theoretical fre-\nquencies are also convoluted with Lorentzian peaks with\nthe same W.RPis always minimized as function of a\nrigid shift of all theoretical frequencies.\nFor the calculations on magnetite we used the vasp\ncode. We used a Monkhorst grid of 6 ×6×2 and an\nenergy cutoff of 400 eV. We used the rotationally in-\nvariantLSDA+ UimplementationbyLichtenstein et al.47\nwith effective on-site Coulomb and exchange parameters:\nU= 4.5 eV48andJ= 0.89 eV for the Fe ions.\nWe used the monoclinic structure as described in\nRefs. 39,49, and calculated the electron density with 56\natoms in the unit cell. In Ref. 39, the charge and mag-\nnetic moment were calculated by integrating the density\nand spin density in a sphere with a radius of 1 ˚A forFe. This radius appears to be chosen such that compara-\nble valueswith neutronand x-raydiffraction experiments\nwere obtained.\nNote, there is no unambiguous way to define these\nradii in systems consisting of two or more atom types.\nTherefore, we checked the correspondence of our results\nto the earlier reported ones and also performed calcula-\ntionswith alargerradiusof1.3 ˚AforFeand0.82 ˚AforO.\nThis is a reasonable choice for Fe mO+\nnclusters since the\noverlapbetween different spheres is minimal, but most of\nthe intra cluster space is covered.\nII. RESULTS AND DISCUSSION\nA. Magnetite\nEven in the bulk, iron oxide is well known for its wide\nvarietyofphasesandtransitions. Magnetite(Fe 3O4), the\nmost stable phase of FemOn, is for example well known\nfor its Verweytransition.1,2Above the transitiontemper-\natureTV, the structure is a cubic inverse spinel. Upon\ncooling below TV, the conductivity decreases by two or-\nders of magnitude due to charge ordering. Furthermore,\nthe structure changes to monoclinic.\nMagnetite has the formal chemical formula\n(Fe3+\nA[Fe2+,Fe3+]BO4) where tetrahedral Asites\nare occupied by Fe3+andBsites contain both divalent\n(Fe2+) and trivalent (Fe3+) iron atoms. Since magnetite\nis a mixed valence system, it is an excellent reference\nsystem for our cluster calculations to determine their\nvalence state and corresponding magnetic moment.\nTABLE I: Spin moments within atomic spheres of 1.3 ˚A for\nthe Fe ions in monoclinic Fe 3O4. For reference the values\nwithin a sphere of 1.0 ˚A are also shown. A and B labels are\nconsistent with Ref. 39.\nSite Spin moment ( µB) Spin moment ( µB)\nRadius sphere 1 .3˚A 1 .0˚A\nFe3+(A) −4.02 −3.78\nFe2+(B1) 3 .69 3 .45\nFe3+(B2) 4 .15 3 .93\nFe3+(B3) 4 .06 3 .84\nFe2+(B4) 3 .64 3 .40\nIn Table I, the spin moments are shown for the dif-\nferent iron ions. The magnetic moments on the Aand\nBsites are antiparallel creating a ferrimagnetic struc-\nture. Within the atomic spheres of 1.3 ˚A the Fe2+and\nFe3+ions have a distinct magnetic moment of 4.0 µB\nand 3.7µBrespectively. Note the difference of 0.3 µBis\nmuch smaller than the 1 µBatomic value and does not\ndepend on the size of the atomic sphere used in the range\nbetween 1.0 and 1.3 ˚A.4\nB. GGA+U\nTo determine the optimal Ueffin comparison to the\nB3LYP functional for the genetic algorithm, we per-\nformed PBE+ Ucalculations on the neutral Fe 3O4clus-\nter. The results for the electronic DOS are shown in\nFig. 1 and compared with the hybrid B3LYP functional.\n/gl507=B==U8nt=/gl72/gl57\n[/gl721[/gl72S=38st/gl3386=\n[/gl721/gl50fS=f8yi=/gl3386=\n[/gl721/gl503S=38UU=/gl3386=/gl56/gl72/gl73/gl73=B=n/gl72/gl57\n/gl507=B==U8nU=/gl72/gl57\n[/gl721[/gl72S=38sf/gl3386=\n[/gl721/gl50fS=f8ys=/gl3386=\n[/gl721/gl503S=38UU=/gl3386=/gl56/gl72/gl73/gl73=B=D/gl72/gl57\n/gl507=B==U8ni=/gl72/gl57\n[/gl721[/gl72S=38sU/gl3386=\n[/gl721/gl50fS=f8ys=/gl3386=\n[/gl721/gl503S=38UU=/gl3386=/gl56/gl72/gl73/gl73=B=3/gl72/gl57H/gl72/gl81/gl86/gl76/gl87/gl92=/gl82/gl73=/gl54/gl87/gl68/gl87/gl72/gl86/gl507=B==U8sn=/gl72/gl57 /gl56/gl72/gl73/gl73=B=f/gl72/gl57\n[/gl721[/gl72S=38sU/gl3386=\n[/gl721/gl50fS=f8yn=/gl3386=\n[/gl721/gl503S=38UU=/gl3386=\n/gl507=B==U8it=/gl72/gl57 /gl56/gl72/gl73/gl73=B=U/gl72/gl57\n[/gl721[/gl72S=38sU/gl3386=\n[/gl721/gl50fS=f8yn=/gl3386=\n[/gl721/gl503S=38UU=/gl3386=\n[/gl721[/gl72S=38sf/gl3386=\n[/gl721/gl50fS=f8yn=/gl3386=\n[/gl721/gl503S=f8oo=/gl3386=dD/gl47/gl60/gl51\n[/gl72=/gl86\n[/gl72=/gl83\n[/gl72=/gl71\n/gl50=/gl86\n/gl50=/gl83M1M−/gl50/gl48/gl50=/gl62/gl72/gl57/gl64=/gl237fU /gl237t8s /gl237s /gl23738s U 38s s\nFIG. 1: (Color online) The density of states for the hybrid\nB3LYP functional and PBE+ Ufor different values of Ueff.\nThe average inter atomic distances are shown on the right,\nwhere Fe-O 1and Fe-O 2refer to the Fe-O distances between\nbridging O atoms (side) and the capping O atom (center),\nrespectively. The mean absolute difference ∆ [Eq. 1] between\nthe PBE+ Uand B3LYP energy levels is also shown and is\nminimal for Ueff= 3 eV, indicating the best match in DOS.\nThe valence states within -4 and 0 eV are formed by\nhybridized orbitals between the dorbitals of iron and\ntheporbitals of oxygen. For increasing U, the majority\nspindorbitals of Fe decrease in energy, whereas HOMO-\nLUMO gap increases. Note that the HOMO-LUMO gap\nof 1.5 eV for Ueff= 4 eV still is 0.9 eV smaller than the\n2.4 eV gap for B3LYP. Furthermore, for Ueff= 2 and\n3 eV the Fe dDOS features are very similar to those of\nthe B3LYP result. To quantify this we also calculated\nthe mean absolute difference ∆ [Eq. 1] for the occupied\nlevels; the results are shown in Fig. 1. ∆ is minimal\nforUeff= 3 eV, indicating the best DOS correspondence\nto B3LYP. We also show the corresponding bonding dis-\ntances within the cluster, where Fe-O 1and Fe-O 2refer\nto the Fe-O distances between bridging O atoms (side)\nand the capping O atom (center), respectively. Note the\ninteratomicdistancesonlychangeverylittlewithincreas-\ningUeff. ForUeff= 3eV,the bindingdistancesarewithin0.01˚A; furthermore, for Ueff= 3 eV and B3LYP the oc-\ncupieddorbitalsofFeareatcomparableenergieswithre-\nspect to the HOMO level. We therefore used Ueff= 3 eV\nfor our genetic algorithm calculations.\nC. Fe 3O0\n4\nAlthough the possible number of isomers increases\nrapidly with cluster size, for small systems such as Fe 3O4\nthe number of possibilities is still small. In Fe 3O4, the Fe\natomscaneitherformatriangleorachain. Forthe trian-\ngular configuration, two isomers are low in energy. The\nfirst isomer consists of a ring like structure where the O\natomsoccupybridgingstatesandoneOatomcapstheFe\ntriangle as is shown in Fig. 2 (a). In the second isomer,\nthe additional O atom is not located above the center\nbut forms an extra bridge between the two ferromagnetic\n(FM) ordered Fe atoms as is shown in Fig. 2 (b).\n3e/gl72/gl57 3e/gl72/gl57 [Sae/gl80/gl72/gl57\n/gl71\n/gl69\n/gl68/gl40/gl81/gl72/gl85/gl74/gl92e/gl62/gl72/gl57/gl64\n335tSS5tpp5ti\n/gl54/gl83/gl76/gl81e/gl80/gl68/gl74/gl81/gl72/gl87/gl76/gl93/gl68/gl87/gl76/gl82/gl81e/gl62/gl541 /gl37/gl643 p a z µ S3 Sp Sa/gl41/gl72/gl22/gl50/gl23/gl19\nFIG. 2: (Color online) The energy as function of spin mag-\nnetization for different neutral Fe 3O4isomers. The geometric\nfigures on the right show the corresponding geometric struc-\nture. O atoms are shown in red, Fe spin up and Fe spin\ndown are indicated with orange (red) and green (blue) colors\n(arrows), respectively. For the lowest magnetic states the rel-\native energy differences are also shown in black. Isomers (a)\n(black line) and (b)(red line) are equally low in energy with\na ferrimagnetic and ferromagnetic ground state, respectiv ely\n(0 eV). The M= 6µBstate of isomer (a)is 14 meV higher\nin energy.\nFigure 2 shows the energy as a function of spin mag-\nnetic moment for the neutral Fe 3O4cluster with four\ndifferent isomers. For all spin magnetizations, the ge-\nometric structure is optimized and shown on the right\nwith its magnetic structure lowest in energy. In Fig. 2\nandthe restofthiswork, Fespin upandFe spindownare\nindicated with orange (red) and green (blue) colors (ar-\nrows), respectively. O atoms are shown in red. For the\nneutral cluster, the two triangular isomers are equally\nlow in energy with two different magnetic configurations.\nThe difference is smaller than 1 meV and therefore be-5\nyond the accuracyofDFT. In isomer (a), as indicated by\nthe black line in Fig. 2, the magnetic ground state corre-\nsponds to ferromagneticalignment between the magnetic\nmoments on the Fe atoms and a total magnetic moment\nof 14µB. The Fe-Fe distances are 2.51 ˚A, the Fe-O dis-\ntances for the bridging O atoms and capping O atom are\n1.84 and 1.99 ˚A, respectively. Aside from the FM ground\nstate, also the ferrimagnetic state with a spin magneti-\nzation of 4 µBis low in energy and only 14 meV higher\nthan the ferromagnetic state. Note we also considered\na noncollinear magnetic state with M= 0µB, but this\nmagnetic configuration did not turn out to be energeti-\ncally stable.\nIsomer(b)is equally low in energy and shown in red\nin Fig. 2. The magnetic ground state corresponds to a\nferrimagnetic alignment where the two ferromagnetically\naligned Fe atoms have Fe-O-Fe angles of approximately\n90◦.\nWe also considered zero point vibrational energies for\nthe three lowest-energy levels. When we include these\ninto our consideration, the ferromagneticstate, indicated\nby the black line, is lowest in energy, and the M= 4µB\nandM= 6µBstatesare17and19meVhigherin energy,\nrespectively.\nD. Fe 3O+\n4\nFor the cation Fe 3O+\n4cluster we also considered ring\nand chain configurations with different oxygen locations.\nFor all four isomers we calculated all possible different\ncollinear magnetic states. Since an antiferromagnetic\n(AFM) triangle is the most simple example of geometri-\ncally frustrated magnetism, we also considered the non-\ncollinear state with M= 0µBwhere all magnetic mo-\nments have 120◦angles with respect to each other. The\nresults are shown in Fig. 3. For the charged Fe 3O+\n4clus-\nter, the isomer with a Fe triangle where the fourth O\natom caps the triangle is, like in the neutral cluster, low-\nestin energy,asisshowninFig.4. Threemagneticstates\nare low in energy: 0, 5 and 15 µB, with the M= 5µB\nstate being lowest in energy, and the non-collinear 0 µB\nand ferromagnetic 15 µBare 20 meV and 58 meV higher\nin energy respectively.\nThe ferrimagnetic state which is lowest in energy, has\na reduced symmetry ( Cv) with respect to the ferromag-\nnetic state ( C3v) and the antiferromagnetic state. This\ncould indicate a Jahn-Teller distortion, but could also\nbe the result of the inability of DFT to correctly model\nthe antiferromagnetic ground state.50,51However, to dis-\ntinguish between these two cases, methods beyond DFT\nsuch as CASPT2 and CCSD(T) are required and there-\nfore beyond the scope of this work. Note that different\nmagnetic states only lead to minor differences in the vi-\nbrational frequencies.\nInterestingly, the typical classical displacement during\na zero-point vibration in these clusters is of the order\nof 0.03˚A. This is of the same order as the typical dif-ytµ8/gl80/gl72/gl57 yp18/gl80/gl72/gl57 18/gl72/gl57\n/gl40/gl81/gl72/gl85/gl74/gl928/gl62/gl72/gl57/gl64\n11.pt1.t1.otSS.ptS.tS.otp\n/gl54/gl83/gl76/gl818/gl80/gl68/gl74/gl81/gl72/gl87/gl76/gl93/gl68/gl87/gl76/gl82/gl818/gl62/gl541 /gl37/gl641 S i t o B SS Si St/gl71\n/gl69\n/gl68/gl41/gl72/gl22/gl50/gl23/gl14\nFIG. 3: (Color online) Energy of the Fe 3O+\n4isomers as func-\ntion of spin magnetization. Figures on the right indicate th e\ncorresponding structure. The isomer lowest in energy (a)is\na Fe triangle with three bridge O atoms and one O atom cap-\npingthetriangle. Forthisisomer, theferrimagnetic 5 µBstate\nis lowest in energy. The antiferromagnetic 0 µBand ferromag-\nnetic 15 µBstate are 20 and 58 meV higher in energy, respec-\ntively. Note the antiferromagnetic 0 µBstate corresponds to a\nnon-collinear orientation with 120◦angles between the spins.\nFIG. 4: (Color online) The neutral (left) and cation (right)\nFe3O4lowest-energy isomers. FespinupandFespindown are\nindicated with orange (red) and green (blue) colors (arrows ),\nrespectively. O atoms are shown in red. The interatomic\ndistances are shown in black. The neutral and cation cluster\nhaveC3vandCvpoint group symmetry, respectively.\nference in inter atomic distances between different mag-\nneticstates. Therefore,this couldleadtointerestingphe-\nnomena in which, for example, there is a strong coupling\nthrough exchange between vibrations and magnetism.\nThe second triangular isomer of Fe 3O+\n4is 154 meV\nhigherinenergyandalsoconsistsofaringstructure. The\nmagnetic state lowest in energy has a magnetic moment\nof 5µB. The Fe-Fe bonding distances are 2.5 and 3.0 ˚A\nbetween the AFM and FM bonds within the structure.\nThe Fe-O distances vary between 1.7 and 1.9 ˚A. The\nisomer has a C2vpoint group symmetry.\nThe third and fourth isomers consist of a linear chain\nof Fe atoms with two O bridging atoms between each Fe\npair. The two planes can be parallel or perpendicular,6\n/gl41/gl72/gl22/gl50/gl23/gl14\n/gl53/gl51 [ I1pbViio /gl80/gl72/gl57 /gl70\n/gl53/gl51 [ I1rnVasb /gl80/gl72/gl57 /gl69\n/gl53/gl51 [ I1sI/gl44/gl53 /gl68/gl69/gl86/gl82/gl85/gl83/gl87/gl76/gl82/gl81 /gl68\n/gl40/gl91/gl83/gl17\n/gl58/gl68/gl89/gl72/gl81/gl88/gl80/gl69/gl72/gl85 /gl62/gl70/gl809a/gl64oII pII iII aIII abII\nFIG. 5: (Color online) The experimental vibration spectra\nof Fe3O+\n4and the calculated isomers lowest in energy. The\nreported energy differences include ZPVE. The Pendry’s reli -\nability factor [Eq. 4] is also shown for each isomer.\nwhere the latter is lower in energy. Both isomers have a\nmagnetic moment of 5 µB.\nIn Fig. 5, both the experimental and calculated vibra-\ntion spectra for the different isomers are shown. The\nexperimental spectrum consists of three peaks at 540,\n610 and 670 cm−1. The best match is given by isomer\n(a)with calculated vibrations at 505, 630 and 660 cm−1\nand a corresponding lowest- RPfactor of 0 .30, indicat-\ning a reasonable match with the experimental spectrum.\nSince isomer (a)is also the lowest in energy, it is identi-\nfied as the experimentally observed structure.\nE. Fe 4O0/+\n5\nFe4O5also consists of a ring structure in which the\nO atoms occupy the bridging sites and one O atom is\nlocated above the center, as is shown in Fig. 6. The clus-\nter has antiferromagnetic order. However, not all Fe-Fe\nbonds are antiferromagnetic, but also two ferromagnet-\nically aligned bonds are present. Therefore, the cluster\nhas noC2vpoint group symmetry but C2, since Fe-Fe\nand Fe-O distances vary between 2.72-2.74 ˚A and 1.79-\n2.33˚A respectively. The magnetic state with four AFM\nFe-Fe bonds is 308 meV higher in energy.\nFor Fe4O+\n5the isomer lowest in energy consists of the\nsame ring structure but is more symmetry broken, since\nthe O atom above the ring is off-center as is shown in\nFig. 6. Therefore the two Fe-Fe distances are 2.69 and\n3.07˚A, the Fe-O distances vary between 1.76 and 2.01 ˚A.\nThe isomer has Cspoint group symmetry. Two Fe 2O2\nsquares are present within the cluster. Isomer (a)has\na magnetic moment of 1 µBdue to ionization. Interest-\ningly, the ionized cluster has a different magnetic ground\nstate with four AFM Fe-Fe bonds opposed to the neutral\ncluster.\nFIG. 6: (color online) The neutral (left) and cation (right)\nFe4O5lowest energy isomers. The neutral cluster has C2sym-\nmetry, whereas the cation cluster has Cssymmetry.\nIn Fig. 7 (b), we also show the vibration spectrum of\nthe ferromagnetic state of this cluster. The Fe-Fe dis-\ntancesareincreasedto2.74and3.11 ˚A, respectively. The\nferromagneticstructureis514meVhigherin energy. The\nvibration spectrum is similar but slightly shifted to the\nblue due to the increased bonding distances.\n/gl53/gl51 - bIisV pvp /gl80/gl72/gl57 /gl71\n/gl53/gl51 - bIttV ptv /gl80/gl72/gl57 /gl70\n/gl53/gl51 - bItbV tsp /gl80/gl72/gl57 /gl69\n/gl53/gl51 - bIpo/gl44/gl53 /gl68/gl69/gl86/gl82/gl85/gl83/gl87/gl76/gl82/gl81 /gl68\n/gl40/gl91/gl83/gl17\n/gl58/gl68/gl89/gl72/gl81/gl88/gl80/gl69/gl72/gl85 /gl62/gl70/gl802s/gl64obb pbb ibb Wbb sbbb sobb/gl41/gl72/gl23/gl50/gl24+\nFIG.7: (Color online)Theexperimentalandcalculated vibr a-\ntion spectra of Fe 4O+\n5. The isomer shown in (a)is both the\nlowest in energy and RP[Eq. 4] and can therefore be iden-\ntified as the experimentally observed geometrical structur e.\nThe reported energy differences include ZPVE.\nThe second isomer, 459meV higher in energy, is shown\nin Fig. 7(c). This cage-like structure has Cvpoint group\nsymmetry and a magnetic moment of 9 µB. Figure 7 (d)\nshows the third isomer which is 494 meV higher in en-\nergy compared to Fig. 7 (a). The isomer has almost no\nsymmetry ( C1), and consists of a ring where one Fe-Fe\nbond has two bridging O atoms. The Fe-Fe binding dis-\ntances vary between 2.62 and 3.13 ˚A. The isomer has a\nmagnetic moment of 1 µB.\nIn the experimental vibration spectrum of Fe 4O+\n5\nshown in Fig. 7, five vibration frequencies can be ob-7\nserved: 450, 615, 760, 810, and 1070 cm−1. The vibra-\ntion at 1070 cm−1can be identified as a shifted vibration\nin the O 2messenger attached to the cluster-messenger\ncomplex and is therefore omitted in the RPcalculation.3\nThe best fit is given by isomer Fig. 7 (a)withRP= 0.42,\nwhich is also the isomer lowest in energy. The calculated\nfrequencies: 479, 630, 637, 772 and 796 cm−1match all\nwithin 30 cm−1to the experimental spectrum. Also, the\nrelative intensities between different vibrations are very\nsimilar. Although the ferromagnetic order increases the\nbinding distances within the cluster, the changes in the\nvibration spectrum of Fig. 7 (b)are small and therefore\nthestructurecorrespondingtoFigs.7 (a)and7(b)canbe\nidentified as the experimentally observed structure and\nthe IR-MPD method is not able to resolve the magnetic\nstate in this case.\nF. Fe 4O0/+\n6\nIn Ref.30, the Fe4O+\n6cluster was already identified as\nthe structure shown in Fig. 8 (b). The reported magnetic\nstructure was ferrimagnetic with a magnetic moment of\n9µB.\n/gl53/gl51 - bIrvesWn /gl80/gl72/gl57 /gl69\n/gl53/gl51 - bIpW/gl44/gl53 /gl68/gl69/gl86/gl82/gl85/gl83/gl87/gl76/gl82/gl81/gl68\n/gl40/gl91/gl83/gl17\n/gl58/gl68/gl89/gl72/gl81/gl88/gl80/gl69/gl72/gl85 /gl62/gl70/gl804s/gl64obb pbb ibb Wbb sbbb sobb/gl41/gl72/gl23/gl50/gl25/gl14\nFIG. 8: (color online) The experimental and calculated vibr a-\ntion spectra of Fe 4O+\n6for both theprevious and newmagnetic\nground state. The vibration frequencies are very similar bu t\ndiffer in absorption intensity. The M= 1µBstate in (a)is\n187 meV lower in energy.\nIn our calculations a magnetic state lower in energy\nwas found for the same geometric structure for both\nFe4O6and Fe 4O+\n6. In this state Fe 4O6and Fe 4O+\n6have a\nmagnetic moment of 0 and 1 µBrespectively as is shown\nin Fig. 9. These structures are 194 and 187 meV lower\nin energy for Fe4O6and Fe4O+\n6in comparison to the\npreviously reported state.30The antiferromagnetic mag-\nnetic ground state of Fe 4O6was also previously reported\nin Ref. 26. For Fe 4O6we also calculated a noncollinear\nstate where all magnetic moments point towardsthe cen-\nter of mass, such state with M= 0µBis 30 meV higher\nin energy compared to the collinear M= 0µBstate.\nFor the neutral cluster, minima in energy are obtained\nforM= 0, 10, 20 µBcorresponding to flips of atomic/gl41/gl72t/gl50z1\n/gl41/gl72t/gl50z[/gl40/gl81/gl72/gl85/gl74/gl924/gl62/gl72/gl57/gl64\n11.i22.iMM.i\n/gl48/gl68/gl74/gl81/gl72/gl87/gl76/gl93/gl68/gl87/gl76/gl82/gl814/gl62/gl541 /gl37/gl641 M t z µ 21 2M 2t 2z 2µ M1/gl41/gl72t/gl50z15[\nFIG. 9: (color online) Energy as function of magnetization o f\nthe neutral Fe 4O6and cationic Fe 4O+\n6clusters. The magnetic\nground state corresponds to a total spin magnetic moment of\nM= 0 and M= 1µBfor Fe 4O6, and Fe 4O+\n6respectively.\nmagnetic moments of 5 µBfor each Fe atom. Note this\nalso matches with an ionic picture in which the Fe atoms\nin Fe4O6have a Fe3+valence state resulting in an atomic\nmagnetic moment of 5 µB. The corresponding structure\nis shown in Fig. 10. In Ref. 30 is mentioned that the\nsymmetry in the M= 10µBstate is reduced from Tdfor\nthe ferromagnetic state to C3v. In this antiferromagnetic\nground state, the neutral cluster has D2dsymmetry. In\nFe4O+\n6the symmetry is reduced even further to Csas is\nshown in Fig. 10.\nFIG. 10: (Color online) The neutral (left) and cation (right )\nFe4O6lowest energy isomers. The neutral cluster has D2d\nsymmetry, whereas the cation cluster has Cssymmetry.\nFigure 8 shows both calculated and experimental spec-\ntra for Fe4O+\n6. The vibration spectra for the two calcu-\nlated magnetic states in Figs. (a)and 8(b)show very\nsimilar behavior. The RPvalues of isomer Fig. 8 (a)\n(0.48) and Fig. 8 (b)(0.39) are both large and indicate a\nbetter match for isomer Fig. 8 (b). Although the spectra\nfor Figs. 8 (a)and 8(b)are very similar, the ferrimag-\nnetic structure has an extra vibration at 720 cm−1with\nsmall IR absorption. Furthermore, around 550 cm−1,\nvibrations differ slightly in frequency. Since the men-\ntioned differences cannot be experimentally resolved, the\nIR-MPD method is unable to resolve between different\nmagnetic states and another type of experiments such\nas Stern-Gerlach deflection is required to determine the\nmagnetic moment.8\nG. Fe 5O0/+\n7\nThe neutral Fe 5O7cluster has a “basket” geometry\nas is shown in Fig. 11. The magnetic ground state is\nferrimagnetic with a total moment of 4 µBdue to the\nodd number of Fe atoms. The cluster has C2vsymmetry.\nFIG. 11: (Color online) The neutral (left) and cation (right )\nFe5O7lowest-energy isomers. The neutral cluster has C2v\nsymmetry, whereas the cation cluster has no symmetry.\nThe cationic structure of Fe 5O+\n7is very different and\nshown in Fig. 11. Like Fe 4O+\n6, it consists of a cage-like\nstructure. The Fe-Fe distances range from 2.7 to 3.1 ˚A.\nExcept for the triple bound O atom, all O atoms form\nbridges between two Fe atoms. The ground state has a\nmagnetic moment of 5 µB. The second isomer is similar\n/gl53/gl51 o vWc[/gl70 3nWvc /gl72/gl57\n/gl53/gl51 o vWrb/gl69 3rv- /gl80/gl72/gl57\n/gl53/gl51 o vWc[/gl68\n/gl40/gl91/gl83/gl17\n/gl58/gl68/gl89/gl72/gl81/gl88/gl80/gl69/gl72/gl85 /gl62/gl70/gl80mn/gl64uvv rvv cvv ]vv nvvv nuvv /gl44/gl53 /gl68/gl69/gl86/gl82/gl85/gl83/gl87/gl76/gl82/gl81/gl41/gl72/gl24/gl50/gl26/gl14\nFIG. 12: (Color online) The experimental and calculated vi-\nbration spectra of Fe 5O+\n7. The reported energy differences\ninclude ZPVE.\nto the neutral ”basket” structure and is 394 meV higher\nin energy as is shown in Fig. 12 (b). The structure has\nCssymmetry and a magnetic moment of 5 µB. However,\nthe atomic spin moments have a different arrangement\nfor the neutral and cationic state.\nThe third isomer is shown in Fig. 12 (c)and is 1.04 eV\nhigher in energy. It contains two triple bonded O atoms\nand is ferrimagnetic with M= 5µB.\nThe experimental vibration spectrum shown in Fig. 12has eight distinct vibrations at 375, 490, 520, 570, 615,\n710, 780, and 830 cm−1which are best resembled by the\nisomer lowest in energy shown in Fig. 12 (a), although\nthegapbetween615and710cm−1seemstobe underesti-\nmated. Note that this also explains the high- RPfactorof\n0.65 for isomer Fig. 12 (a). Similar to Fe 4O+\n5and Fe 4O+\n6\nthe absorption intensities of vibrations in the range of\n300-500 cm−1are systematically underestimated. The\nindividualvibrationsofisomerFig. 12 (a)areallin agree-\nment within 35 cm−1. Although isomer Fig. 12 (b)has a\nlowerRP= 0.43, the energy difference of 407 meV with\nisomer Fig. 12 (a)is large and isomer Fig. 12 (b)has a\nvibration at 450 cm−1which is not present in the exper-\nimental spectrum and lacks the experimental 375 cm−1\nvibration. Therefore, isomer Fig. 12 (a)can be identified\nas the most probable ground state.\nH. Fe 6O+\n8\nThe isomer lowest in energy found for Fe6O+\n8is shown\ninFig.13andhas Cssymmetrywherethereflectionplane\nis located through Fe atoms 1, 3, and 6. The magnetic\nmoment of this isomer is 1 µB.\nFIG. 13: (Color online) The cation Fe 6O+\n8isomer lowest in\nenergy. The cluster has Cssymmetry.\nThesecondisomerlowinenergyisshowninFig.14 (b).\nIn this isomer no symmetry is present. Compared to the\nlowest found isomer in Fig. 14 (a)it is 413 meV higher in\nenergy and also has a magnetic moment of 1 µB.\nFigure 14 (c)shows the third isomer, which is a dis-\ntorted octahedral of Fe atoms in which the O atoms cap\nthe Fe triangles. The structure is slightly distorted due\nto the AFM order between spins, which lead to slightly\naltered Fe-Fe distances. This isomer is 483 meV higher\nin energy than isomer Fig. 14 (a).\nFigure 14 also shows the corresponding vibration spec-\ntra of the mentioned isomers and the experimental spec-\ntrum. The experimental spectrum has vibrations at 392,\n420, 500, 730 and 763 cm−1. Note that none of the\nprovided isomers match the experimental vibration spec-\ntrum completely. This is also shown by the large- RP\nvalues of 0.56-0.61 for all calculated isomers. The isomer\nlowest in energy Fig. 14 (a)is the best match since it also\nhasvibrationsat 420and500cm−1, but the vibrationsat9\n/gl53/gl51 o a1r]/gl70 eb-u /gl80/gl72/gl57\n/gl53/gl51 o a1[v/gl69 ebvu /gl80/gl72/gl57\n/gl53/gl51 o a1r[/gl68\n/gl40/gl91/gl83/gl17\n/gl58/gl68/gl89/gl72/gl81/gl88/gl80/gl69/gl72/gl85 /gl62/gl70/gl806v/gl64naa baa [aa -aa vaaa vnaa /gl44/gl53 /gl68/gl69/gl86/gl82/gl85/gl83/gl87/gl76/gl82/gl81/gl41/gl72/gl25/gl50/gl27/gl14\nFIG. 14: (color online) The experimental and calculated vi-\nbration spectra of Fe 6O+\n8. The isomer shown in (a)is the low-\nest in energy. The reported energy differences include ZPVE.\n804 and 825 are considerably shifted with respect to 730\nand 763 cm−1. Furthermore, the vibrations at 640, 671,\nand 713 cm−1are not present in the experimental spec-\ntrum. The vibration spectra shown in Figs. 14 (b)and\n14(c)fit even worse. Therefore, we can not successfully\nidentify the Fe 6O+\n8structure.\nNote that our genetic algorithm implementation only\nuses geometry optimization at the DFT level. At cluster\nsizes of Fe 6O+\n8and larger, preselection using empirical\npotentials instead of immediate geometry optimization\nusing DFT might be more efficient in generating possible\nisomers.\nI. Electronic structure\nIn the bulk, iron-oxide materials have many different\ncrystal structures such as hematite, wustite, and mag-\nnetite with all corresponding different electronic struc-\ntures. While in hematite only trivalent Fe3+is present,\nthe mixed valence state (Fe3+\nA[Fe2+,Fe3+]BO4) in mag-\nnetite leads to interesting physical phenomena such as\nferrimagnetic ordering between the sublattices Aand\nBand the Verwey transition in which orbital ordering\nleads to a first-order phase transition in the electrical\nconductivity.1,2\nIn clusters, stoichiometries corresponding to both\nhematite (Fe4O6) and magnetite (Fe3O4, Fe6O8) and\nother combinations (Fe 4O5, Fe5O7) occur. We therefore\nexpect divalent and trivalent Fe cations to be present in\nthe reported clusters. There is no unique method to de-\ntermine the valence state in materials consisting of mul-\ntiple types of elements. We therefore compare both the\nlocal magnetic moments and the local density of states\n(LDOS) for our cluster calculations with bulk magnetite\nresults shown in Section IIA. Since the Fe2+and Fe3+\nfeatures in the LDOS are very similar for different clustersizes, we show the LDOS of Fe 4O+\n5which contains both\nFe2+and Fe3+in Fig. 15. The LDOS for other cluster\nsizes can be found in the Appendix.\n/gl41/gl720/gl502 /gl47/gl39/gl50/gl54 /gl76/gl81/gl87s4/gl39/gl50/gl54 /gl55/gl82/gl87s4/gl39/gl50/gl54/gl41/gl724/gl86\n/gl41/gl724/gl83\n/gl41/gl724/gl71\n/gl504/gl86\n/gl504/gl83\n/gl88/gl83\n/gl71/gl82/gl90/gl81\n2udu2wdw21d12\n/gl41/gl72u\n/gl41/gl72w\n/gl41/gl721\n/gl41/gl720\n/gl40o/gl40/gl43/gl50/gl48/gl504/gl62/gl72/gl57/gl64/gl237- /gl2373 /gl2370 /gl237w d w 0 3\nFIG. 15: (Color online) The total, integrated and local den-\nsity of states of the Fe atoms for the Fe 4O+\n5cluster. The\ntrivalent Fe(1), Fe(2) and Fe(3) all show 3 dlevels at -6 eV\nand small hybridization with O. The divalent Fe(4), however ,\nshows strong hybridization and a single level at E HOMO.\nTable II shows the local spin moments of the clus-\nters: Fe 3O+\n4, Fe4O+\n5, Fe4O+\n6, Fe5O+\n7and Fe 6O+\n8. For\nFe3O+\n4all three Fe atoms have a similar spin moment\nwithin 0.04 µB. A comparison with magnetite suggests\nall Fe atomsare trivalent. This agreeswith an ionic bond\nmodel. Furthermore, this is confirmed by the integrated\nand local density of states shown in Appendix A. The 3 d\npeaksaround-6eVcorrespondto15electrons,indicating\nthe hybridization between Fe and O is small. Note that,\nthe central oxygen atoms O(4) and O(7) are partially\nspin polarized.\nFor Fe4O+\n5, the spin moment of Fe(4) is 0.5 µBlower\nthan the other Fe atoms, indicating three trivalent and a\nsingle divalent atom. The difference is also in agreement\nwith the magnetite results. The Fe(4) also breaks the C2\nsymmetry as is shown in Fig. 6. The local (LDOS) and\nintegrated density of states are shown in Fig. 15. Note\nthat all Fe3+have 3dpeaks around −6 eV and small\nhybridization with O is present, similar to the Fe 3O+\n4\ncluster. The LDOS of the divalent Fe(4) atom however\nshows strong hybridization with O and a single minority\nlevel at E HOMO.\nWhereasFe 4O6onlycontains trivalent Fe,26for Fe4O+\n6\nthis is no longer the case due to ionization. As can be\nseen from Table II, three trivalent Fe atoms are present,\ntogether with a single Fe4+atom. The spin moment is10\nreduced with respect to Fe3+, consistent with a higher\noxidation state than Fe3+.\nIn Fe5O+\n7, only trivalent Fe atoms are present, con-\nsistent with an ionic model and the ionized state of the\ncluster. Fe 6O+\n8, on the other hand, is again a mixed\nvalence cluster where the magnetic moment of Fe(4) is\n0.4µBlower than the other Fe atoms, indicating Fe(4)\nis divalent. This is also consistent with the LDOS shown\nin Appendix A.\nFigure 16 shows the density of states for the different\ncationicclustersandmagnetite. Thecalculatedbandgap\nof 0.2 eV in magnetite is considerably smaller than for\nthereportedclusters: around3eVforFe 3O+\n4andslightly\nsmaller for Fe 4O+\n5and Fe 4O+\n6. Furthermore, whereas\nmagnetite has a t2gorbital of Fe2+just below the Fermi\nenergy,39in the reported clusters Fe4O+\n5and Fe6O+\n8have\nasimilarlevelduetoadivalentFeatom. Notethatthe3 d\norbitals of Fe3+in the clusters are located around 5.5 eV\nbelow the HOMO level, which is 2 eV higher in energy\ncompared to magnetite.\n/gl39/gl72/gl81/gl86/gl76/gl87/gl92n/gl82/gl73n/gl54/gl87/gl68/gl87/gl72/gl86/gl41/gl72g/gl50-6\n/gl41/gl72M/gl50E6/gl41/gl72n/gl86\n/gl41/gl72n/gl83\n/gl41/gl72n/gl71\n/gl50n/gl86\n/gl50n/gl83n\n/gl41/gl723/gl50g6\n/gl41/gl723/gl50M6\n/gl41/gl724/gl5036\n/gl48/gl68/gl74/gl81/gl72/gl87/gl76/gl87/gl72\n/gl408/gl40/gl43/gl50/gl48/gl50n/gl62/gl72/gl57/gl64/gl237- /gl237g /gl2373 /gl237d 7 d 3 g\nFIG. 16: (Color online) The density of states for Fe xO+\nyclus-\nters. For these calculations a smearing of 0.15 eV was used\nfor convenience of the reader. The HOMO level is located at\n0 eV and the small occupation above the HOMO level is due\nto smearing.III. CONCLUSION\nIn this work, we have studied the geometric, elec-\ntronic and magnetic structure of Fe xO+\nyclusters using\ndensity functional theory. For Fe 3O4we compared bind-\ning distances and electronic structure between the hybrid\nB3LYP functional, and different Ueffin the PBE+ Ufor-\nmalism. We found the best match for Ueff= 3 eV. Using\nthe PBE+ Uformalism and a genetic algorithm, many\npossible isomers were considered. For isomers low in en-\nergy, all different magnetic configurations were further\ngeometrically optimized. Finally, for the cationic clus-\nters we calculated the vibration spectra and compared\nthem with experiments to identify the geometric struc-\nture of Fe 3O+\n4, Fe4O+\n5, Fe4O+\n6, Fe5O+\n7and Fe 6O+\n8. All\ncationic clusters with an even number of Fe atoms have\na small magnetic moment of 1 µBdue to ionization. Fur-\nthermore, comparison with bulk magnetite reveals that\nFe4O+\n5, Fe4O+\n6and Fe 6O+\n8are mixed valence clusters.\nIn contrast, in Fe3O+\n4and Fe5O+\n7all Fe are found to be\ntrivalent.\nIV. ACKNOWLEDGEMENTS\nThe work is supported by European Research Council\n(ERC) Advanced Grant No. 338957 FEMTO/NANO.\n1E. J. W. Verwey, Nature 144, 327 (1939).2F. Walz, J. Phys.: Condens. Matter 14, R285 (2002).11\n3L. Andrews, G. V. Chertihin, A. Ricca, and C. W.\nBauschlicher, J. Am. Chem. Soc. 118, 467 (1996).\n4G. V. Chertihin, W. Saffel, J. T. Yustein, L. Andrews,\nM. Neurock, A. Ricca, and C. W. Bauschlicher, J. Phys.\nChem.100, 5261 (1996).\n5S. Laurent, D. Forge, M. Port, A. Roch, C. Robic, L. Van-\nder Elst, and R. N. Muller, Chem. Rev. 108, 2064 (2008).\n6N. M. Reilly, J. U. Reveles, G. E. Johnson, S. N. Khanna,\nand A. W. Castleman, J. Phys. Chem. A 111, 4158 (2007).\n7L. S. Wang, H. Wu, and S. R. Desai, Phys. Rev. Lett. 76,\n4853 (1996).\n8D. Schr¨ oder, P. Jackson, and H. Schwarz, Eur. J. Inorg.\nChem.2000, 1171 (2000).\n9A. Erlebach, H. D. Kurland, J. Grabow, F. A. M¨ uller, and\nM. Sierka, Nanoscale 7, 2960 (2015).\n10B. V. Reddy, F. Rasouli, M. R. Hajaligol, and S. N.\nKhanna, Fuel 83, 1537 (2004).\n11B. V. Reddy and S. N. Khanna, Phys. Rev. Lett. 93,\n068301 (2004).\n12A. Fiedler, D. Schroeder, S. Shaik, and H. Schwarz, J. Am.\nChem. Soc. 116, 10734 (1994).\n13K. Ohshimo, T. Komukai, R. Moriyama, and F. Misaizu,\nJ. Phys. Chem. A 118, 3899 (2014).\n14S. Yin, W. Xue, X. L. Ding, W. G. Wang, S. G. He, and\nM. F. Ge, Int. J. Mass Spectrom. 281, 72 (2009).\n15K. Palot´ as, A. N. Andriotis, and A. Lappas, Phys. Rev. B\n81, 075403 (2010).\n16Q. Sun, Q. Wang, K. Parlinski, J. Z. Yu, Y. Hashi, X. G.\nGong, and Y. Kawazoe, Phys. Rev. B 61, 5781 (2000).\n17Q. Wang, Q. Sun, M. Sakurai, J. Z. Yu, B. L. Gu,\nK. Sumiyama, and Y. Kawazoe, Phys. Rev. B 59, 12672\n(1999).\n18J. Kortus and M. R. Pederson, Phys. Rev. B 62, 5755\n(2000).\n19Q. Sun, B. V. Reddy, M. Marquez, P. Jena, C. Gonzalez,\nand Q. Wang, J. Phys. Chem. C 111, 4159 (2007).\n20S. L´ opez, A. H. Romero, J. Mej´ ıa-L´ opez, J. Mazo-Zuluaga,\nand J. Restrepo, Phys. Rev. B 80, 085107 (2009).\n21X. L. Ding, W. Xue, Y. P. Ma, Z. C. Wang, and S. G. He,\nJ. Chem. Phys. 130, 014303 (2009).\n22N. O. Jones, B. V. Reddy, F. Rasouli, and S. N. Khanna,\nPhys. Rev. B 72, 165411 (2005).\n23H. Shiroishi, T. Oda, I. Hamada, and N. Fujima, Eur.\nPhys. J. D 24, 85 (2003).\n24Q. Sun, M. Sakurai, Q. Wang, J. Z. Yu, G. H. Wang,\nK. Sumiyama, and Y. Kawazoe, Phys. Rev. B 62, 8500\n(2000).\n25Z. Cao, M. Duran, and M. Sol` a, J. Chem. Soc., Faraday\nTrans.94, 2877 (1998).\n26A. Erlebach, C. H¨ uhn, R. Jana, and M. Sierka, Phys.\nChem. Chem. Phys. 16, 26421 (2014).\n27R. L. Johnston, Dalton Trans. 2003, 4193 (2003).\n28M. Haertelt, A. Fielicke, G. Meijer, K. Kwapien, M. Sierka,\nand J. Sauer, Phys. Chem. Chem. Phys. 14, 2849 (2012).\n29H. J. Zhai, J. D¨ obler, J. Sauer, and L. S. Wang, J. Am.\nChem. Soc. 129, 13270 (2007).\n30A.Kirilyuk, A.Fielicke, K.Demyk,G. vonHelden, G. Mei-\njer, and T. Rasing, Phys. Rev. B 82, 020405 (2010).\n31G. Kresse and J. Furthm¨ uller, Phys. Rev. B 54, 11169\n(1996).32P. E. Bl¨ ochl, Phys. Rev. B 50, 17953 (1994).\n33G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).\n34J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.\nLett.77, 3865 (1996).\n35V. I. Anisimov and Y. Izyumov, Electronic Structure of\nStrongly Correlated Materials (Springer-Verlag Berlin Hei-\ndelberg, 2010).\n36V. I. Anisimov, F. Aryasetiawan, and A. I. Lichtenstein, J.\nPhys.: Condens. Matter 9, 767 (1997).\n37A. M. Burow, T. Wende, M. Sierka, R. Wodarczyk,\nJ. Sauer, P. Claes, L. Jiang, G. Meijer, P. Lievens, and\nK. R. Asmis, Phys. Chem. Chem. Phys. 13, 19393 (2011).\n38S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J.\nHumphreys, and A. P. Sutton, Phys. Rev. B 57, 1505\n(1998).\n39H. T. Jeng, G. Y. Guo, and D. J. Huang, Phys. Rev. Lett.\n93, 156403 (2004).\n40A. D. Becke, J. Chem. Phys. 98, 1372 (1993).\n41G. Makov and M. C. Payne, Phys. Rev. B 51, 4014 (1995).\n42J. Neugebauer and M. Scheffler, Phys. Rev. B 46, 16067\n(1992).\n43L. Fan and T. Ziegler, J. Chem. Phys. 96, 9005 (1992).\n44D. Porezag and M. R. Pederson, Phys. Rev. B 54, 7830\n(1996).\n45J. B. Pendry, J. Phys. C 13, 937 (1980).\n46M. Rossi, V. Blum, P. Kupser, G. Von Helden, F. Bierau,\nK. Pagel, G. Meijer, and M. Scheffler, J. Phys. Chem. Lett.\n1, 3465 (2010).\n47A. I. Liechtenstein, V. I. Anisimov, and J. Zaanen, Phys.\nRev. B52, 5467 (1995).\n48V.I.Anisimov, I.S.Elfimov, N.Hamada, andK.Terakura,\nPhys. Rev. B 54, 4387 (1996).\n49J. P. Wright, J. P. Attfield, and P. G. Radaelli, Phys. Rev.\nLett.87, 266401 (2001).\n50C. J. Cramer andD. G. Truhlar, Phys. Chem. Chem. Phys.\n11, 10757 (2009).\n51M. Reiher, Faraday Discuss. 135, 97 (2007).\n52S. H. Vosko, L. Wilk, and M. Nusair, Can. J. Phys. 58,\n1200 (1980).\n53In particular, we use B3LYP with the VWN3 functional as\ndefined in Ref. 52.\nAppendix A: Local DOS\nIn this appendix we show the integrated and local\nDOS of the clusters Fe3O+\n4, Fe4O+\n6, Fe5O+\n7, Fe6O+\n8,\nand magnetite. Figures 17, 18, 19, and 20 show the\ntotal, integrated and local density of states of Fe3O+\n4,\nFe4O+\n6, Fe5O+\n7, and Fe 6O+\n8, respectively. Of these clus-\nters, Fe3O+\n4and Fe5O+\n7are pure trivalent and the LDOS\ncontains 3 dpeaks at -6 eV and small hybridization be-\ntween Fe and O. Fe 4O+\n6contains a single tetravalent Fe\natom, with a similar LDOS compared to Fe3+. The ion-\nized electron is not removed from the 3d levels at -6 eV,\nbut from the hybridized levels with oxygen, as can be\nseen from the integrated density of states. Fe 4O+\n5and\nFe6O+\n8contain a single divalent Fe atom, which has a\ndistinct LDOS, in which there are no peaks around -6 eV12\nbut strong spin polarized hybridization with oxygen and\nasingleoccupiedminoritylevelattheHOMOlevel. Even\nin bulk magnetite, as is shown in Fig. 21, the same fea-\ntures between divalent and trivalent Fe atoms exist.\n/gl41/gl72u/gl50w /gl47/gl39/gl50/gl54 /gl76/gl81/gl87a3/gl39/gl50/gl54 /gl55/gl82/gl87/gl68/gl793/gl39/gl50/gl54/gl41/gl723/gl86\n/gl41/gl723/gl83\n/gl41/gl723/gl71\n/gl503/gl86\n/gl503/gl83\n/gl88/gl83\n/gl71/gl82/gl90/gl81\n5psp5dsd5\n/gl41/gl72p\n/gl41/gl72d\n/gl41/gl72u\n/gl40o/gl40/gl43/gl50/gl48/gl503/gl62/gl72/gl57/gl64/gl237ps /gl2372 /gl2371 /gl237w /gl237d s d w 1\nFIG.17: (Color online)Thetotal, integratedandlocal dens ity\nof states of the Fe 3O+\n4cluster.\n/gl41/gl720/gl502 /gl47/gl39/gl50/gl54 /gl76/gl81/gl87s4/gl39/gl50/gl54 /gl55/gl82/gl87s4/gl39/gl50/gl54/gl41/gl724/gl86\n/gl41/gl724/gl83\n/gl41/gl724/gl71\n/gl504/gl86\n/gl504/gl83\n/gl88/gl83\n/gl71/gl82/gl90/gl81\nudu5wdw51d15\n/gl41/gl72u\n/gl41/gl72w\n/gl41/gl721\n/gl41/gl720\n/gl40o/gl40/gl43/gl50/gl48/gl504/gl62/gl72/gl57/gl64/gl237E /gl2372 /gl2370 /gl237w d w 0 2\nFIG. 18: (Color online) The total, integrated, and local den -\nsity of states of the Fe 4O+\n6cluster. Fe(1) is tetravalent as is\nshown in Table I.\n/gl47/gl39/gl50/gl54/gl41/gl722/gl504 /gl76/gl81/gl87sO/gl39/gl50/gl54 /gl55/gl82/gl87sO/gl39/gl50/gl54/gl41/gl72O/gl86\n/gl41/gl72O/gl83\n/gl41/gl72O/gl71\n/gl50O/gl86\n/gl50O/gl83\n/gl88/gl83\n/gl71/gl82/gl90/gl81\nudwd1d0d2d\n/gl41/gl72u\n/gl41/gl72w\n/gl41/gl721\n/gl41/gl720\n/gl41/gl722\n/gl40o/gl40/gl43/gl50/gl48/gl50O/gl62/gl72/gl57/gl64/gl237E /gl2373 /gl2370 /gl237w d w 0 3\nFIG. 19: (Color online) The total, integrated, and local den -\nsity of states of the Fe 5O+\n7cluster.13\nTABLE II: The spin moment for Fe xO+\nyclusters. The atom numbers correspond to the atom numbers sh own in Figures 4,6,10,\n11, and 13. The spin moment is calculated using atomic sphere s of 1.3 and 0.82 ˚A for Fe and O, respectively.\nCluster Spin moment [ µB]\n1 2 3 4 5 6 7 8\nFe3O+\n4 Fe −3.84 3 .88 3 .88\nO 0 .56 0 .00 0 .00 0 .22\nFe4O+\n5 Fe 3 .89 −3.84 3 .89 −3.40\nO −0.05 0 .13 0 .20 −0.05 0 .20\nFe4O+\n6 Fe −3.22 3 .85 3 .85 −3.79\nO 0 .01 0 .54 0 .01 −0.25 0 .00 0 .00\nFe5O+\n7 Fe 3 .85 3 .87 3 .89 −3.83 −3.80\nO 0 .01 0 .10 0 .03 0 .51 −0.09 0 .05 0 .12\nFe6O+\n8 Fe 3 .80 −3.84 3 .85 −3.47 −3.84 3 .88\nO 0 .01 0 .51 0 .01 0 .01 −0.10 0 .17 −0.10 0 .0114\n/gl41/gl723/gl505 /gl47/gl39/gl50/gl54 /gl76/gl81/gl87s6/gl39/gl50/gl54 /gl55/gl82/gl87s6/gl39/gl50/gl54/gl41/gl726/gl86\n/gl41/gl726/gl83\n/gl41/gl726/gl71\n/gl506/gl86\n/gl506/gl83\n/gl88/gl83\n/gl71/gl82/gl90/gl81\nudwd1d0d2d\n/gl41/gl72u\n/gl41/gl72w\n/gl41/gl721\n/gl41/gl720\n/gl41/gl722\n/gl41/gl723\n/gl40o/gl40/gl43/gl50/gl48/gl506/gl62/gl72/gl57/gl64/gl2375 /gl2373 /gl2370 /gl237w d w 0 3\nFIG. 20: (Color online) The total, integrated, and local den -\nsity of states of the Fe 6O+\n8cluster. All Fe atoms are trivalent\nexcept for Fe(4), which is divalent./gl48/gl68/gl74/gl81/gl72/gl87/gl76/gl87/gl72/gl47/gl39/gl50/gl54 /gl55/gl82/gl873g/gl39/gl50/gl54/gl41/gl722.S/gl36T\n/gl41/gl72).S/gl37AT\n/gl41/gl722.S/gl37)T\n/gl41/gl722.S/gl372T\n/gl41/gl72).S/gl37BT\n/gl40F/gl40/gl41g/gl62/gl72/gl57/gl64/gl237- /gl2374 /gl237B /gl237) ( ) B 4\nFIG. 21: (Color online) The total and local density of states\nof the different Fe atoms in magnetite. The numbering is\nconsistent with Table I. Fe2+and Fe3+have a similar LDOS\nto clusters although the symmetry is very different." }, { "title": "2111.15142v1.First_and_second_order_magnetic_anisotropy_and_damping_of_europium_iron_garnet_under_high_strain.pdf", "content": "1 \n First and second order magnetic anisotropy and damping of \neuropium iron garnet under high strain \n \nVíctor H. Ortiz1, Bassim Arkook1, Junxue Li1, Mohammed Aldosary1, Mason Biggerstaff1, Wei \nYuan1, Chad Warren2, Yasuhiro Kodera2, Javier E. Garay2, Igor Barsukov1*, and Jing Shi1* \n \n1Department of Physics and Astronomy, University of California, Riverside, CA 92521, \nUSA \n2 Department of Mechanical and Aerospace Engineering, University of California, San \nDiego, CA 92093, USA \n \n \nUnderstanding and tailoring static and dynamic properties of magnetic insulator thin films \nis important for spintronic device applications. Here, we grow atomically flat epitaxial europium \niron garnet (EuIG) thin films by pulsed laser deposition on (111) -oriented garnet sub strates with a \nrange of lattice parameters. By controlling the lattice mismatch between EuIG and the substrates, \nwe tune the strain in EuIG films from compressive to tensile regime, which is characterized by X -\nray diffraction. Using ferromagnetic resonance , we find that in addition to the first -order \nperpendicular magnetic anisotropy which depends linearly on the strain, there is a significant \nsecond -order one that has a quadratic strain dependence. Inhomogeneous linewidth of the \nferromagnetic resonance inc reases notably with increasing strain, while the Gilbert damping \nparameter remains nearly constant (≈ 2× 10-2). The se results provide valuable insight into the spin \ndynamics in ferrimagnetic insulators and useful guidance for material synthesis and engineer ing \nof next -generation spintronics applications. \n \n \n*: Corresponding authors: Igor Barsukov ( igorb@ucr.edu ) and Jing Shi ( jing.shi@ucr.edu ) 2 \n Ferrimagnetic insulators (FMIs) have played an important role in uncovering a series of \nnovel spintronic effects such as spin Seebeck effect (SSE) and spin Hall magnetoresistance (SMR). \nIn addition, FMI thin films have proved to be an excellent source of proximity -induced \nferromagnetism in adjacent layers (e.g., heavy metals [1], graphene [1] and topological \ninsulators [2]) and of pure spin currents [3–6]. FMIs have also been shown to be a superb medium \nfor magnon spin currents with a long decay length [7,8]. Among FMIs, rare earth iron garnets \n(REIGs) in particular have a plethora of desirable properties for practical applications: high Curie \ntemperature (T c > 550 K), strong chemical stability, and relatively large band gaps (~ 2.8 eV). \nCompared to other magnetic materials, REIGs are distinct owing to their magnetoelastic \neffect with the magnetostriction coefficient ranging from -8.5×106 to +21 ×106 at room \ntemperature [9] and up to two orders of magnitude increases at low temperatures [10]. This unique \nfeature allows for tailoring ma gnetic anisotropy in REIG thin films via growth, for example, by \nmeans of controlling lattice mismatch with substrates, film thickness, oxygen pressure, and \nchemical substitution. In thin films, the magnetization usually prefers to be in the film plane due \nto magnetic shape anisotropy; however, the competing perpendicular magnetic anisotropy (PMA) \ncan be introduced by utilizing magneto -crystalline anisotropy or interfacial strain, both of which \nhave been demonstrated through epitaxial growth [11–14]. In the study of Tb 3Fe5O12 (TbIG) and \nEu3Fe5O12 (EuIG) thin films, the PMA field H2ꓕ was found to be as high as 7 T under interfacial \nstrain [11], much stronger than the demagnetizing field. While using strain is proven to be an \neffective way of manipulating magn etic anisotropy, it often comes at a cost of increasing magnetic \ninhomogeneity and damping of thin films [15,16]. \nIn this work, we investigate the effect of strain on magnetic properties of (111) -oriented \nEuIG thin films for the following reasons: (1) The spin dynamics in EuIG bulk crystals is \nparticularly interesting but has not been studied thoroughly in the thin film form. Compared to \nother REIGs, the Eu3+ ions occupying the dodecahedral sites (c -site) should have the J = 0 ground \nstate according to the Hund’s rules, which do not contribute to the total magnetic moment; \ntherefore, EuIG thin films can potentially have a ferromagnetic resonance (FMR) linewidt h as \nnarrow as that of Y 3Fe5O12 (YIG) [17,18] or Lu 3Fe5O12 (LuIG) [19]. In EuIG crystals, a very \nnarrow linewidth (< 1 Oe) [20] was indeed observed at low temperatures, but it showed a nearly \ntwo orders of magnitude increase at high temperatures, which ra ises fundamental questions \nregarding the damping mechanism responsible for this precipitous change. (2) Although it has 3 \n been shown that the uniaxial anisotropy can be controlled by moderate strain for different substrate \norientations and even in polycrysta lline form [21], the emergence of the higher -order anisotropy \nat larger strain, despite its technological significance, has remained elusive. \nWe grow EuIG films by pulsed laser deposition (PLD) from a target densified by powders \nsynthesized using the meth od described previously [22]. The films are deposited on (111) -oriented \nGd3Sc2Ga3O12 (GSGG), Nd 3Ga5O12 (NGG), Gd 2.6Ca0.4Ga4.1Mg 0.25Zr0.65O12 (SGGG), \nY3Sc2Ga3O12 (YSGG), Gd3Ga5O12 (GGG), Tb 3Ga5O12 (TGG) and Y 3Al5O12 (YAG) single crystal \nsubstrates, with the lattice mismatch 𝜂=𝑎𝑠𝑢𝑏𝑠𝑡𝑟𝑎𝑡𝑒 −𝑎𝐸𝑢𝐼𝐺\n𝑎𝐸𝑢𝐼𝐺 (where 𝑎 represents the lattice parameter \nof the referred material) ranging from +0.45% (GSGG) to -3.95% (YAG) in the decreasing order \n(see Table I). After the standard solvent cleaning process, the substrates are annealed at 220 °C \ninside the PLD chamber with the base pressure lower than 10-6 Torr for 5 hours prior to deposition . \nThen the temperature is increased to ~ 600 °C in the atmosphere of 1.5 mT orr oxygen mixed with \n12% (wt.) ozone for 30 minutes. A 248 nm KrF excimer pulsed laser is used to ablate the target \nwith a power of 156 mJ and a repetition rate of 1 Hz. We crystalize the films by ex situ annealing \nat 800 °C for 200 s in a steady flow of oxygen using rapid thermal annealing (RTA) . \nReflection high energy electron diffraction (RHEED) is used to evaluate the crystalline \nstructural properties of the EuIG films grown on various substrates (Fig. 1a). Immediately after \nthe deposition, RHEED dis plays the absence of any crystalline order. After ex situ rapid thermal \nannealing, all EuIG films turn into single crystals. We carry out atomic force microscopy (AFM) \non all samples and find that they show atomic flatness and good uniformity with root -mean-square \n(RMS) roughness < 2 Å (Fig. 1b). In addition, we perform X -ray diffraction (XRD) on all samples \nusing a Rigaku SmartLab with Cu K α radiation with a Ni filter and Ge(220) mirror as \nmonochromators, at room temperature in 0.002° steps over the 2 range from 10° to 90° [23]. In \na representative XRD spectrum (Fig. 1c), two (444) Bragg peaks are present, one from the 50 nm \nthick EuIG film and the other from the YSGG substrate, which confirms the epitaxial growth and \nsingle crystal structure of the fi lm without evidence of any secondary phases. Other REIG films \ngrown under similar conditions , i.e., by PLD in oxygen mixed with ozone at ~600 °C during \nfollowed by RTA, have shown no observable interdiffusion across the interface from high \nresolution trans mission electron microscopy and energy dispersive X -ray spectroscopy (Fig. S1 , \n[24]). The EuIG Bragg peak ( a0 = 12.497 Å) is shifted with respect to the expected peak position \nof unstrained bulk crystal, indicating a change in the EuIG lattice parameter pe rpendicular to the 4 \n surface ( aꓕ). For the example shown in Fig. 1c, the EuIG (444) peak shifts to left with respect to \nits bulk value, indicating an out -of-plane tensile strain and therefore an in -plane compressive strain \nin the EuIG lattice. \n \nA common app roach for inferring the in -plane strain ε|| of thin films from the standard −2 \nXRD measurements involves the following equation [23], \n \n𝜀∥= −𝑐11+2 𝑐12+4 𝑐44\n2𝑐11+4 𝑐12−4 𝑐44 𝜀⊥, with 𝜀⊥=𝑎⊥−𝑎𝑜\n𝑎𝑜, (1) \n \nwhere a0 is the lattice parameter of the bulk material, and aꓕ can be calculated using 𝑎⊥=\n𝑑ℎ𝑘𝑙√ℎ2+𝑘2+𝑙2 from the interplanar distance 𝑑ℎ𝑘𝑙 obtained from the XRD data (Fig. S 2, [25]), \nand cij are the elastic stiffness constants of the crystal which in most cases can be found in the \nliterature [9]. However, due to the wide range of strain values studied in this work and the \npossibility that the films may contain different amounts of crystalline defects, we perform \nreciprocal space mapping (RSM) measurements on a subset of our EuIG samples (Fig. S 3, [26]) \nand compared the measured in -plane lattice parameters with the calculated ones using Eq. 1. We \nobserve that the average in -plane strain s measured by RSM has a systematic difference of 40% \nfrom the calculated values based on the elastic properties (Fig. S 4, [26]). Given this nearly constant \nfactor for all measured films, we find that the elastic stiffness constants of our EuIG films may \ndeviate from the literature reported bulk values , possibly due to stochiometric deviations or slight \nunit cell distortion in thin films . Here we adopt the reported lattice parameter value ( a0 = 12.497 \nÅ) as the reference due to the difficulty of grow ing sufficiently thick, unstrained EuIG films usin g \nPLD . \nIn the thickness -tuned magnetic anisotropy study [11], the anisotropy field in REIG films is \nfound to be proportional to η/(t+t o), which was attributed to the relaxation of strain as the film \nthickness t increases. Here in EuIG samples with small lattice mismatch η (e.g., NGG/EuIG), the \nstrain is mostly preserved in 50 nm thick films (pseudomorphic regime), whereas for larger η (e.g., \nYAG/EuIG ), the lattice parameter of EuIG films shows nearly complete structural relaxation to \nthe bulk value. For this reason, in the samples with larger η (YAG = -3.95 %, GSGG = 0.45%), \nwe grow thinner EuIG films (20 nm) in order to retain a larger in -plane strain (compressive for 5 \n YAG, tensile for GSGG). For EuIG films gr own on TGG and GGG substrates, the paramagnetic \nbackground of the substrates is too large to obtain a reliable magnetic moment measurement of the \nEuIG films; therefore, the results of thinner films on these two substrates are not included in this \nstudy. \nRoom-temperature magnetic hysteresis curves for YSGG/EuIG sample are shown in Fig. 1d \nwith the magnetic field applied parallel and perpendicular to the film [26]. The saturation field for \nthe out -of-plane loop (~1100 Oe) is clearly larger than that for the i n-plane loop, indicating that \nthe magnetization prefers to lie in the film plane. Moreover, since the demagnetizing field 4π Ms \n(≈ 920 Oe) is less than the saturation field in the out -of-plane loop (Fig. S 5, [27]), it suggests the \npresence of additional easy -plane anisotropy result ing from the magnetoelastic effect due to \ninterfacial strain. As shown in this example, we can qualitatively track the evolution of the \nmagnetic anisotropy in samples with different strains. However, this approach cannot provide a \nquantitative description when high -order anisotropy contributions are involved. \nTo quantitatively determine magnetic anisotropy in all EuIG films, we perform polar angle \n(H)-dependent FMR measurements using an X -band microwave cavity with f requency f = 9.32 \nGHz and field modulation. The samples are rotated from H = 0° to H = 180° in 10° steps, where \nH = 90° corresponds to the field parallel to the sample plane (Fig. 2a). The spectra at Η = 0° for \nall samples are displayed in Fig. 2b and show a single resonance peak which can be well fitted by \na Lorentzian derivative. Despite different strains in all s amples, the resonance field Hres is lower \nfor the in -plane direction ( H = 90°) than for the out -of-plane direction ( H = 0°). A quick \ninspection reveals that the out -of-plane Hres shifts to larger values as η increases in the positive \ndirection (e.g., fro m YAG/EuIG to GSGG/EuIG), corresponding to stronger easy -plane \nanisotropy. Furthermore, the Hres values at θΗ = 0° show a large spread among the samples. Fig. 2c \nshows a comparison of FMR spectra at different polar angles between two representative samples: \nNGG/EuIG (small η) and YAG/EuIG (large η). \nFigs. 3a -c show Hres vs. θH for three representative EuIG films . To evaluate magnetic \nanisotropy, we fit the data using the Smit -Beljers formalism by considering the first -order \n−𝐾1cos2𝜃 and the second -order −1\n2𝐾2cos4𝜃 uniaxial anisotropy energy terms [28]. From this \nfitting, we extract the parameters 4𝜋𝑀𝑒𝑓𝑓=4𝜋𝑀𝑠− 2𝐾1\n𝑀𝑠= 4𝜋𝑀𝑠- 𝐻2⊥ and 𝐻4⊥= 2𝐾2\n𝑀𝑠 (see Table \nI), her e 𝐻2⊥ and 𝐻4⊥ being the first- and second - order anisotropy fields, respectively , and favoring 6 \n out-of-plane (in -plane) orientation of magnetization when they are positive (negative). The \nspectroscopic g-factor is treated as a fitted parameter which is found as a nearly constant , g = \n1.40 (Fig. S 6, [28]), in accordance to the previous results obtained by Miyadai [31]. In Figs. 3d \nand 3e, we present 𝐻2⊥ and 𝐻4⊥ as functions of the measured out -of-plane strain 𝜀⊥ and in -plane \nstrain 𝜀∥. Clear ly, the magnitude of 4𝜋𝑀𝑒𝑓𝑓 is greater than the demagnetizing field for EuIG \n4𝜋𝑀𝑠=920 𝑂𝑒; therefore, 𝐻2⊥ is negative for all samples, i.e., favoring the in -plane orientation. \nAs shown in Fig. 3d, |𝐻2⊥| increases linearly with increasing in -plane strain η. This is consistent \nwith the magnetoelastic effect in (111) -oriented EuIG films [9]. As briefly discussed earlier, due \nto the constant scaling factor between the calculated and measured 𝜀∥, we rewrite t he \nmagnetoelastic contribution to the first -order perpendicular anisotropy as −9𝛯\n3𝑀𝑠𝜀⊥, with the \nparameter 𝛯 containing the information related to the magnetoelastic constant λ111 and elastic \nstiffness cii. We fit the magnetoelastic equation in Ref. [11] using the parameter 𝛯 and obtain 𝛯=\n−(7.06±0.95)×104 𝑑𝑦𝑛𝑒\n𝑐𝑚2 from the slope. On the other hand, based on the reported literature \nvalues ( 𝜆111=+1.8×10−6, c11 = 25.10 ×1011 dyne/cm2, c12 = 10.70 ×1011 𝑑𝑦𝑛𝑒\n𝑐𝑚2, c44 = 7.62 ×1011 \n𝑑𝑦𝑛𝑒\n𝑐𝑚2) [10], we obtain 𝛯𝑙𝑖𝑡=−6.12×104 𝑑𝑦𝑛𝑒\n𝑐𝑚2. This result suggests that even though the actual \nelastic properties of our EuIG films may be different from the ones reported in for EuIG crystals \ndue to the thin film unit cell distortion (Table S1 , [32]), the pertaining parameter 𝛯 appears to be \nrelatively insensitive to variations of stoichiometry . The intercept of the straight -line fit should \ngive the magneto -crystalline anisotropy coefficient of EuIG Kc. We find Kc = (+62.76 ± 0.18 ) × \n103 erg/cm3, which is differ ent from the previously reported values for EuIG bulk crystals in both \nthe magnitude and sign ( Kc = -38 × 103 erg/cm3) [31]. Similar growth -modified magneto -\ncrystalline anisotropy was observed in EuIG films grown with relatively lo w temperatures \n(requiring post -deposition annealing to crystalize) [10]. In the absence of interfacial interdiffusion, \nthe anomalous anisotropy may be related to partial deviation from the chemical ordering of the \ngarnet structure [31]. \n By comparing the first - and second -order anisotropy fields 𝐻2⊥ and 𝐻4⊥ vs. 𝜀∥ plotted in \nFigs. 3d and 3e, we find that the former dominates over the entire range of 𝜀∥ (except for \nYAG/EuIG). In contrast to the linear dependence for 𝐻2⊥, 𝐻4⊥ can be fitted well with a quadratic \n𝜀∥ dependence , which is not surprising for materials with large magnetostriction constants (such 7 \n as EuIG) under large strains. For relatively small 𝜀∥, the linear strain term in the magnetic \nanisotropy energy dictates . For large 𝜀∥, higher -order strain terms may not be neglected. By \nincluding the ( 𝜀∥cos2θ)2 term, we obtain excellent fitting to the FMR data, indicating that the \nsecond -order expansion in 𝜀∥ is adequate. In contrast to 𝐻2⊥, 𝐻4⊥ is always positive, thus favoring \nout-of-plane magnetization orientation. It is worth pointing out that for YAG and TGG, the \nmagnitude of the 𝐻2⊥ becomes comparable with that of the 𝐻4⊥, but the sign differ s. Comparison \nof 𝐻4⊥ with 4𝜋𝑀𝑒𝑓𝑓 reveals that a coexistence (bi -stable) magnetic state can be realized when \n𝐻4⊥>4𝜋𝑀𝑒𝑓𝑓 [31, 33 -35]. The results are summarized in Table I. \nThe above magnetic anisotropy energy analysis only deals with the polar angle dependence , \nbut in principle, it can also vary in the film plane and therefore depend on the azimuthal angle. To \nunderstand the latter, w e perform azimuthal angle dependent FMR measurements on all samples. \nWe indeed observe a six -fold in -plane anisotropy in Hres due to the crystalline symmetry of EuIG \n(111). However, the amplitude of the six -fold Hres variation is less than 15 Oe, about two orders \nof magnitude smaller than the average value of Hres for most samples, thus we omit the in-plane \nanisotropy in our analysis. \nBesides the Hres information, t he FMR spectra in Fig. 2 c reveal s significant variations in \nFMR linewidth, which contains information of magnetic inhomogeneity and Gilbert damping. To \ninvestigate these properties systematically, we perform broad -band (up to 15 GHz) FMR \nmeasurements with m agnetic field applied in the film plane, using a coplanar waveguide setup. \nFrom the frequency dependence of Hres, we obtain 4𝜋𝑀𝑒𝑓𝑓 and g independently via fitting the data \nwith the Kittel equation. These values agree very well with those previously found from the polar \nangle dependence. We plot the half width at half maximum, ∆𝐻, as a function of frequency f in \nFig. 4a. While ∆𝐻 varies significantly across the samples, the data for each sample fall \napproximately on a straight line and the slope of ∆𝐻 vs. 𝑓 appears to be visibly close to each other. \nFor a quantitative evaluation of ∆𝐻, we consider the following contributions: the Gilbert damping \n∆𝐻𝐺𝑖𝑙𝑏𝑒𝑟𝑡 , two -magnon scatt ering ∆𝐻𝑇𝑀𝑆, and the inhomogeneous linewidth ∆𝐻0 [36], \n \n∆𝐻=∆𝐻𝐺𝑖𝑙𝑏𝑒𝑟𝑡 +∆𝐻𝑇𝑀𝑆 +∆𝐻0 . (3) \n 8 \n The Gilbert term, ∆𝐻𝐺𝑖𝑙𝑏𝑒𝑟𝑡 =2𝜋𝛼𝑓\n|𝛾|, depends linearly on f, where α is the Gilbert damping \nparameter; the two -magnon term is described through ∆𝐻𝑇𝑀𝑆 =𝛤0𝑎𝑟𝑐𝑠𝑖𝑛 √√𝑓2+(𝑓𝑜\n2)2\n−𝑓𝑜\n2\n√𝑓2+(𝑓𝑜\n2)2\n+𝑓𝑜\n2 [37], \nwhere 𝛤0 denotes the magnitude of the two -magnon scattering, f0 = 2γMeff; and ∆𝐻0, the \ninhomogeneous linewidth w hich is frequency independent. \nBy fitting Eq. (3) to the linewidth data, we obtain quantitative information on magnetic \ndamping through the Gilbert parameter and two -magnon scattering magnitude as well as the \nmagnetic inhomogeneity [39–40]. In Fig. 4a, the overall linear behavior for all samples is an \nindication of a relatively small two -magnon scattering contribution ∆𝐻𝑇𝑀𝑆 which therefore may \nbe disregarded in the fitting process. Figs. 4b and 4c show both ∆𝐻0 and α vs. 𝜀∥. It is cl ear that \nfour of the samples with the smallest ∆𝐻0 (~ 10 Oe) are those with relatively low in -plane strain \n(|𝜀∥|<0.30% ). In the meantime, the XRD spectra of these samples show fringes characteristic of \nwell conformed crystal planes (Fig. S 2), and moreover, the RSM plots (Fig. S 3) reveal a uniform \nstrain distribution in the films [41]. On the compressive strain side, ∆𝐻0 increases steeply to 400 \nOe at 𝜀∥ ~ -0.40 %, and their XRD spectra show no fringes and the RSM graphs indicate non-\nuniform strain relaxation in the samples (Figs. S2 and S3 ). In sharp contrast to the ∆𝐻0 trend, the \nGilbert damping α remains about 2 ×10-2 over the entire range of 𝜀∥, sugges ting that the intrinsic \nmagneti c damping of EuIG films is nearly unaffected by the inhomogeneity. In fact, the magnitude \nof α is significantly larger than that of YIG [17,18] or LuIG films [19], which is somewhat \nunexpected for Eu3+ in EuIG with J = 0. A possible reason for this enhance d damping is that other \nvalence states of Eu such as Eu2+ (J =7/2) may be present, which leads to non -zero magnetic \nmoments of Eu ions in the EuIG lattice and thus results in a larger damping constant, common to \nother REIG with non -zero 4f -moments [42]. The X -ray photoelectron spectroscopy data taken on \nYSGG(111)/EuIG(50 nm) (Fig. S7 , [43]) indicates such a possibility. While the FMR linewidth \npresents large variations across the sample set, we have identified that the non-uniform strain \nrelaxation process caused by large lattice mismatch with the substrate is a main source of the \ninhomogeneity linewidth ∆𝐻0, but it does not affect the Gilbert damp ing α. The results raise \ninteresting questions on the mechanisms of intrinsic damping and the origin of magnetic \ninhomogeneity in EuIG thin films , both of which warrant further investigations. 9 \n In summary, we find that uniaxial magnetic anisotropy in PLD -grown EuIG(111) thin films \ncan be tuned over a wide range via magnetostriction and lattice -mismatch induced strain. The first -\norder anisotropy field depends linearly on the strain and the second order anisotropy field has a \nquadratic dependence. While non -uniform strain relaxation significantly increases the magnetic \ninhomogeneity, the Gilbert damping remains nearly constant over a wide range of in -plane strain. \nThe results demonstrate broad tunab ility of magnetic properties in REIG films and provide \nguidance for implementation of EuIG for spintronic applications. Further studies to elucidate the \nrole of Eu2+ sites in magnetic damping are called upon. \n \nWe thank Dong Yan and Daniel Borchardt for the ir technical assistance. This work was supported \nas part of the SHINES, an Energy Frontier Research Center funded by the US Department of \nEnergy, Office of Science, Basic Energy Sciences under Award No. SC0012670. J.S. \nacknowledges support by DOE BES Award No. DE -FG02 -07ER46351 and I.B. acknowledges \nsupport by the National Science Foundation under grant number NSF -ECCS -1810541. \n 10 \n References \n \n[1] Z. Wang, C. Tang, R. Sachs, Y. Barlas, and J. Shi, Proximity -Induced Ferromagnetism in \nGraphene Revealed by the Anomalous Hall Effect , Phys. Rev. Lett. 114, 016603 (2015). \n[2] Z. Jiang, C. -Z. Chang, C. Tang, P. Wei, J. S. Moodera, and J. Shi, Independent Tuning of \nElectronic Properties and Induced Ferromagnetism in Topological Insulators with \nHeterostructure Approach , Nano Lett. 15, 5835 (2015). \n[3] K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. \nUmezawa, H. Kawai, G. E. W. Bauer, S. Maekawa, and E. Saitoh, Spin Seebeck Insulator , \nNat. Mater. 9, 894 (2010). \n[4] Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. \nKawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Transmission of Electrical \nSignals by Spin -Wave Interconversion in a Magnetic Insulator , Nature 464, 262 (201 0). \n[5] J. Li, Y. Xu, M. Aldosary, C. Tang, Z. Lin, S. Zhang, R. Lake, and J. Shi, Observation of \nMagnon -Mediated Current Drag in Pt/Yttrium Iron Garnet/Pt(Ta) Trilayers , Nat. \nCommun. 7, 10858 (2016). \n[6] V. H. Ortiz, M. J. Gomez, Y. Liu, M. Aldosary, J. S hi, and R. B. Wilson, Ultrafast \nMeasurements of the Interfacial Spin Seebeck Effect in Au and Rare -Earth Iron -Garnet \nBilayers , Phys. Rev. Mater. 5, 074401 (2021). \n[7] L. J. Cornelissen, J. Liu, R. A. Duine, J. Ben Youssef, and B. J. van Wees, Long -Distance \nTransport of Magnon Spin Information in a Magnetic Insulator at Room Temperature , \nNat. Phys. 11, 1022 (2015). \n[8] B. L. Giles, Z. Yang, J. S. Jamison, and R. C. Myers, Long -Range Pure Magnon Spin \nDiffusion Observed in a Nonlocal Spin -Seebeck Geometry , Phy s. Rev. B 92, 224415 \n(2015). \n[9] S. Iida, Magnetostriction Constants of Rare Earth Iron Garnets , J. Phys. Soc. Japan 22, \n1201 (1967). \n[10] P. Hansen, Magnetic Anisotropy and Magnetostriction in Garnets , in Rendiconti Della \nScuola Internazionale Di Fisica “ Enrico Fermi” (1978), pp. 56 –133. \n[11] V. H. Ortiz, M. Aldosary, J. Li, Y. Xu, M. I. Lohmann, P. Sellappan, Y. Kodera, J. E. \nGaray, and J. Shi, Systematic Control of Strain -Induced Perpendicular Magnetic 11 \n Anisotropy in Epitaxial Europium and Terbium Iron Ga rnet Thin Films , APL Mater. 6, \n121113 (2018). \n[12] M. Kubota, K. Shibuya, Y. Tokunaga, F. Kagawa, A. Tsukazaki, Y. Tokura, and M. \nKawasaki, Systematic Control of Stress -Induced Anisotropy in Pseudomorphic Iron \nGarnet Thin Films , J. Magn. Magn. Mater. 339, 63 (2013). \n[13] E. R. Rosenberg, L. Beran, C. O. Avci, C. Zeledon, B. Song, C. Gonzalez -Fuentes, J. \nMendil, P. Gambardella, M. Veis, C. Garcia, G. S. D. Beach, and C. A. Ross, Magnetism \nand Spin Transport in Rare -Earth -Rich Epitaxial Terbium and Europium I ron Garnet \nFilms , Phys. Rev. Mater. 2, 094405 (2018). \n[14] Y. Krockenberger, K. S. Yun, T. Hatano, S. Arisawa, M. Kawasaki, and Y. Tokura, \nLayer -by-Layer Growth and Magnetic Properties of Y 3Fe5O12 Thin Films on Gd 3Ga5O12, \nJ. Appl. Phys. 106, 108 (2009). \n[15] H. Wang, C. Du, P. C. Hammel, and F. Yang, Strain -Tunable Magnetocrystalline \nAnisotropy in Epitaxial Y 3Fe5O12 Thin Films , Phys. Rev. B 89, 134404 (2014). \n[16] B. Bhoi, B. Kim, Y. Kim, M. Kim, J. Lee, and S. -K. Kim, Stress -Induced Magnetic \nProp erties of PLD -Grown High -Quality Ultrathin YIG Films , J. Appl. Phys. 123, 203902 \n(2018). \n[17] C. Tang, M. Aldosary, Z. Jiang, H. Chang, B. Madon, K. Chan, M. Wu, J. E. Garay, and J. \nShi, Exquisite Growth Control and Magnetic Properties of Yttrium Iron Garn et Thin \nFilms , Appl. Phys. Lett. 108, (2016). \n[18] Y. Sun, Y. Y. Song, H. Chang, M. Kabatek, M. Jantz, W. Schneider, M. Wu, H. \nSchultheiss, and A. Hoffmann, Growth and Ferromagnetic Resonance Properties of \nNanometer -Thick Yttrium Iron Garnet Films , Appl. P hys. Lett. 101, (2012). \n[19] C. L. Jermain, H. Paik, S. V. Aradhya, R. A. Buhrman, D. G. Schlom, and D. C. Ralph, \nLow-Damping Sub -10-Nm Thin Films of Lutetium Iron Garnet Grown by Molecular -Beam \nEpitaxy , Appl. Phys. Lett. 109, (2016). \n[20] R. C. LeCraw, W. G. Nilsen, J. P. Remeika, and J. H. Van Vleck, Ferromagnetic \nRelaxation in Europium Iron Garnet , Phys. Rev. Lett. 11, 490 (1963). \n[21] J. J. Bauer, E. R. Rosenberg, and C. A. Ross, Perpendicular Magnetic Anisotropy and \nSpin Mixing Conductance in Polycrystalline Europium Iron Garnet Thin Films , Appl. \nPhys. Lett. 114, 052403 (2019). 12 \n [22] P. Sellappan, C. Tang, J. Shi, and J. E. Garay, An Integrated Approach to Doped Thin \nFilms with Strain Tunable Magnetic Anisotropy: Powder Synthesis, Target Prepara tion \nand Pulsed Laser Deposition of Bi:YIG , 3831 , 1 (2016). \n[23] E. Anastassakis, Strained Superlattices and Heterostructures: Elastic Considerations , J. \nAppl. Phys. 68, 4561 (1990). \n[24] See Supplemental Information figure S1 at http://placeholder.html fo r the HRTEM image \nand EDS mapping. \n[25] See Supplemental Information figure S2 at http://placeholder.html for the θ−2θ HRXRD \nof the samples. \n[26] See Supplemental Information figure S3 and S 4 at http://placeholder.html for the XRD \nand strain analysis of the samples. \n[27] See Supplemental Information figure S 5 at http://placeholder.html for the M vs H \nhysteresis loops of the samples. \n[28] I. Barsukov, Y. Fu, A. M. Gonçalves, M. Spasova, M. Farle , L. C. Sampaio, R. E. Arias, \nand I. N. Krivorotov, Field -Dependent Perpendicular Magnetic Anisotropy in CoFeB \nThin Films , Appl. Phys. Lett. 105, 152403 (2014). \n[29] See Supplemental Information figure S6 at http://placeholder.html for the g -factor \ncompari son of the samples. \n[30] T. Miyadai, Ferrimagnetic Resonance in Europium -Iron Garnet , J. Phys. Soc. Japan 15, \n2205 (1960). \n[31] F. B. Hagedorn, Annealing Behavior and Temperature Dependence of the Growth -\nInduced Magnetic Anisotropy in Epitaxial Sm -YIGG , J. Appl. Phys. 45, 3123 (1974). \n[32] See Supplemental Information Table S1 at http://placeholder.html for the calculated \nparameters of the rhombohedral distorted unit cell . \n[33] Y. Fu, I. Barsukov, J. Li, A. M. Gonçalves, C. C. Kuo, M. Farle, and I. N. Krivorotov, \nTemperature Dependence of Perpendicular Magnetic Anisotropy in CoFeB Thin Films , \nAppl. Phys. Lett. 108, (2016). \n[34] R. Skomski, H. P. Oepen, and J. Kirschner, Unidirect ional Anisotropy in Ultrathin \nTransition - Metal Films , Phys. Rev. B 58, 138 (1998). \n[35] J. M. Shaw, H. T. Nembach, M. Weiler, T. J. Silva, M. Schoen, J. Z. Sun, and D. C. \nWorledge, Perpendicular Magnetic Anisotropy and Easy Cone State in 13 \n Ta/Co 60Fe20B20/MgO, IEEE Magn. Lett. 6, 1 (2015). \n[36] I. Barsukov, P. Landeros, R. Meckenstock, J. Lindner, D. Spoddig, Z. A. Li, B. Krumme, \nH. Wende, D. L. Mills, and M. Farle, Tuning Magnetic Relaxation by Oblique Deposition , \nPhys. Rev. B - Condens . Matter Mater. Phys. 85, 1 (2012). \n[37] J. Lindner, K. Lenz, E. Kosubek, K. Baberschke, D. Spoddig, R. Meckenstock, J. Pelzl, Z. \nFrait, and L. Mills, Non-Gilbert -Type Damping of the Magnetic Relaxation in Ultrathin \nFerromagnets: Importance of Magnon -Magno n Scattering , Phys. Rev. B - Condens. \nMatter Mater. Phys. 68, 6 (2003). \n[38] A. Navabi, Y. Liu, P. Upadhyaya, K. Murata, F. Ebrahimi, G. Yu, B. Ma, Y. Rao, M. \nYazdani, M. Montazeri, L. Pan, I. N. Krivorotov, I. Barsukov, Q. Yang, P. Khalili Amiri, \nY. Tserk ovnyak, and K. L. Wang, Control of Spin -Wave Damping in YIG Using Spin \nCurrents from Topological Insulators , Phys. Rev. Appl. 11, 1 (2019). \n[39] A. Etesamirad, R. Rodriguez, J. Bocanegra, R. Verba, J. Katine, I. N. Krivorotov, V. \nTyberkevych, B. Ivanov, an d I. Barsukov, Controlling Magnon Interaction by a \nNanoscale Switch , ACS Appl. Mater. Interfaces 13, 20288 (2021). \n[40] I. Barsukov, H. K. Lee, A. A. Jara, Y. J. Chen, A. M. Gonçalves, C. Sha, J. A. Katine, R. \nE. Arias, B. A. Ivanov, and I. N. Krivorotov, Giant Nonlinear Damping in Nanoscale \nFerromagnets , Sci. Adv. 5, 1 (2019). \n[41] X. Guo, A. H. Tavakoli, S. Sutton, R. K. Kukkadapu, L. Qi, A. Lanzirotti, M. Newville, \nM. Asta, and A. Navrotsky, Cerium Substitution in Yttrium Iron Garnet: Valence State, \nStructure, and Energetics , (2013). \n[42] C. Tang, P. Sellappan, Y. Liu, Y. Xu, J. E. Garay, and J. Shi, Anomalous Hall Hysteresis \nin Tm 3Fe5O12/Pt with Strain -Induced Perpendicular Magnetic Anisotropy , Phys. Rev. B \n94, 140403 (2016). \n[43] See Supplemental Information figure S7 at http://placeholder.html for the XPS spectra and \nanalysis. \n \n 14 \n Figures \n \n \n \n \n \n \n \n \nFigure SEQ Figure \\* ARABIC 1 : Structural and magnetic property characterization of \nEuIG 50 nm film grown on YSGG(111) substrate. (a) Reflection high energy electron \ndiffraction (RHEED) pattern along the direction, displaying single crystal structure after \nrapid thermal anneal ing process. (b) 2 mm 2 mm atomic force microscope (AFM) surface \nmorphology scan, demonstrating a root -mean -square (RMS) roughness of 1.7 Å. (c) \nIntensity semi -log plot of \n - 2\n XRD scan. The dashed line corresponds to the XRD peak \nfor bulk EuIG. (d) Mag netization hysteresis loops for field out -of-plane and in -plane \ndirections. Figure 1: Structural and magnetic property characterization of EuIG 50 nm film grown on \nTGG(111) substrate . (a) Reflection high energy electron diffraction (RHEED ) pattern along the \n⟨112⟩ direction, displaying single crystal structure after rapid thermal annealing process. (b) 5 mm \n 5 mm atomic force microscope ( AFM ) surface morphology scan, demonstrating a root-mean -\nsquare (RMS) roughness of 1.8 Å. (c) Intensity semi -log plot of - 2 XRD scan. The dashed line \ncorresponds to the XRD peak for bulk EuIG. (d) Magnetization hysteresis loops for field out -of-\nplane and in -plane directions. 15 \n \n \nFigure 2 Polar angle dependent ferromagnetic resonance (FMR). (a) Coordinate system used for \nthe FMR measurement. (b) Room temperature FMR derivative absorption spectra for θH = 0° (out -\nof-plane configuration) for EuIG on different (111) substrates. (c) FMR derivative absorption \nspectra for 50 nm EuIG grown on NGG(111) ( 𝜀∥ ≈ 0) and 20 nm EuIG on YAG(111) (𝜀∥< 0) with \npolar angle θH ranging from 0° (out -of-plane) to 90° ( in-plane) at 300 K, where 𝜀∥ is in-plane strain \nbetween the EuIG film and substrate. \n \n \n \n16 \n \nFigure 3 Polar angle dependent ferromagnetic resonance field Hres for (a) tensile in -plane strain \n(𝜀∥ > 0), (b) in -plane strain close to zero ( 𝜀∥ ≈ 0), and (c) compressive in -plane strain ( 𝜀∥ < 0). Solid \ncurves represent the best fitting results. In -plane strain dependence of the anisotropy fields H 2ꓕ (d) \nand H 4ꓕ (e). \n \n \n \n \n \n17 \n \nFigure 4 FMR linewidth and magnetic damping of EuIG films as a function of in -plane strain. (a) \nHalf width at half maximum ∆𝐻 vs. frequency f for EuIG films grown on different substrates, with \nthe corresponding fitting according to Eq. (3). In -plane strain depen dence of inhomogeneous \nlinewidth ΔH0 (b) and Gilbert parameter α (c). \n \n \n \n \n \n \n \n \n \n \n \n \n \n18 \n \n \nSubstrate asubstrate \n(Å) η \n(%) t \n(nm) 𝜀∥ (%) 𝜀⊥ (%) g H2ꓕ \n(Oe) H4ꓕ \n(Oe) α (×10-2) ΔHo \n(Oe) Γo (Oe) \nGSGG 12.554 0.45 50 0.34 -0.16 1.40 -1394.2 \n± 44.9 339.79 \n± 6.59 2.46 ± \n0.03 21.4 ± \n1.3 2.61 \n 25 0.46 -0.21 1.41 -1543.6 \n± 39.7 709.47 \n± 27.5 1.58 ± \n0.06 10.2 ± \n1.7 6.05 \nNGG 12.508 0.06 50 0.12 -0.06 1.38 -1224.4 \n± 5.7 18.34 ± \n0.05 2.41 \n±0.01 8.9 ± \n0.7 0.20 \nSGGG 12.480 -\n0.14 50 -0.13 0.06 1.40 -909.6 \n± 15.2 164.8 ± \n1.36 2.13 ± \n0.01 5.6 ± \n0.4 0.50 \nYSGG 12.426 -\n0.57 50 -0.27 0.12 1.37 -709.4 \n± 22.0 377.3 ± \n5.09 2.47 ± \n0.03 9.9 ± \n1.8 2.47 \nGGG 12.383 -\n0.92 50 -0.45 0.21 1.38 -1015.0 \n± 81.3 887.2 ± \n37.27 2.20 ± \n0.14 412.2 \n± 8.4 3.35 \nTGG 12.355 -\n1.14 50 -0.38 0.18 1.38 -393.4 \n± 53.6 245.0 ± \n10.00 2.29 ± \n0.20 253.4 \n± 11.8 0.20 \nYAG 12.004 -\n3.95 20 -0.42 0.20 1.37 -36.8 ± \n47.1 424.8 ± \n20.91 1.86 ± \n0.20 217.0 \n± 22.6 0.20 \n \nTable 1 Structural and magnetic parameters for the EuIG thin films grown on different substrates. " }, { "title": "2005.03965v3.Sublattice_magnetizations_of_ultrathin_ferrimagnetic_lamellar_nanostructures_between_cobalt_leads.pdf", "content": "arXiv:2005.03965v3 [cond-mat.mes-hall] 12 Apr 2023SPIN\nVol. 1, No. 1 (2022) 1–11\n©World Scientific Publishing Company\nSublattice magnetizations of ultrathin ferrimagnetic lam ellar\nnanostructures between cobalt leads\nVinod Ashokan*\nDepartment of Physics,\nDr. B. R. Ambedkar National Institute of Technology,\nJalandhar (Punjab) 144 027, India\nashokanv@nitj.ac.in\nA. Khater\nDepartment of Physics, Le Mans University, 72085 Le Mans, Fr ance;\nDepartment of Theoretical Physics, Jan Dlugosz University, Czestochowa, Poland\nM. Abou Ghantous\nScience Department, American University of Technology,\nFidar Campus, Halat, Lebanon\nIn this work we model the salient magnetic properties of the alloy lame llar ferrimagnetic nanos-\ntructures [ Co1−cGdc]ℓ′[Co]ℓ[Co1−cGdc]ℓ′between Cosemi-infinite leads. We have employed the\nIsing spin effective field theory (EFT) to compute the reliable magnet ic exchange constants for\nthe pure cobalt JCo−Coand gadolinium JGd−Gdmaterials, in complete agreement with their\nexperimental data. The sublattice magnetizations of the CoandGdsites on the individual hcp\natomic (0001) planes of the Co−Gdlayered nanostructures are computed for each plane and\ncorresponding sites, by using the combined EFT and mean field theor y (MFT) spin methods.\nThe sublattice magnetizations, effective site magnetic moments, an d ferrimagnetic compensa-\ntion characteristics for the individual hcp atomic planes of the embe dded nanostructures, are\ncomputed as a function of temperature, and for various stable eu tectic concentrations in the\nrangec≤0.5. The theoretical results for the sublattice magnetizations and the local magnetic\nvariablesof these ultrathin ferrimagneticlamellar nanostructured systems, between cobalt leads,\nare necessary for the study of their magnonic transport proper ties, and eventually their spin-\ntronic dynamic computations. The method developed in this work is ge neral and can be applied\nto comparable magnetic systems nanostructured with other mate rials.\nKeywords : effective field theory; mean field theory; sublattice magnetization ; exchange constant;\ncobalt-gadolinium alloy; ferrimagnetic nanojunction.\n1. Introduction\nThe multilayered lamellar nano-magnetic nanos-\ntructurehas madea tremendousprogress in prepar-\ning and analyzing the physical properties of nano-\nmagnetic layered nanostructures and magneticnano-junctions. These systems have a panoply of\nindustrial and technological applications in the ar-\neas of spin wave magnonics1;2;3;4, and spintron-\nics5;6;7. However, the study of nano-magnetic\nlamellar multilayered nanostructure and nanojunc-\n∗Corresponding author\n12Vinod Ashokan; A. Khater and M. Abou Ghantous\ntion composite of rare earth-transition metal alloy\nsystems are still in its infancy. A fundamental and\nintriguing interest associated with such rare earth-\ntransition metal systems is to understand the phe-\nnomenawhichmay ariseduetothedecreaseoftheir\nsize when surface and quantum mechanical effects\ncome into play. The atomic interfaces and sublat-\ntice magnetization in these nanostructures are be-\ncome critical components for the physics of embed-\nded nano-junctions.\nThe nano-magnetic properties of ferrimag-\nnetic alloy nanostructure of rare earth-transition\nmetal has been reported in the literature\n8;9;10;11;12;13;14. The nano fabrication tech-\nnique has made the experimental progress to real-\nize the thin multilayered nanostructures with novel\nphysical properties and promising applications in\nmagnonic devices. In this work we study in par-\nticular the Co-Gd rare earth-transition metal alloy\nsystem; the transition metal Coand rare earth Gd\nare ferromagnetic with their Curie temperatures of\n119.4 and 25.2 meV respectively. Nano-magnetic\nnanojunctions build from Co-Gd alloys in diverse\nmultilayer formats can present hence very useful\nproperties for technological applications at room\ntemperature. In this respect, the bulk Co 1−cGdcal-\nloy materials with different alloy concentrations c,\nhave been studied intensively in the past for diverse\napplications in sensors,magneto optical devices and\nmagnetic storage elements15;16.\nA greater understanding of lamellar Co/Gd\nmultilayers nanostructures has been achieved\n9;11;12. It is noted that when we form an amor-\nphous alloy Co1−cGdcin these systems there is a\nstrong asymmetric spontaneous diffusion of Cointo\ntheGdplane occurs, and interfaces for various con-\ncentration. The experimental techniques made pos-\nsible to control the interdiffusion for some stable\neutectic concentration c≤0.5 while preserving the\nferrimagnetic structure of the multilayer systems.\nThe properties of multilayers presenting alloy in-\nterfaces, with few lamellar atomic planes thick sig-\nnificantly depends on the degree of material inter-\ndiffusion9;11. Such interdiffusion may play crucial\nrole in determining the nano-magnetic properties\nof the multilayers systems8, because the individ-\nual planes are having different exchange couplings\nat their interfaces from the bulk. This has also been\nreported earlier by model calculations which show\nthat nano-atomic scale magnetic alloyed interfaces\ncan significantly modify the nano-magnetic proper-ties of multilayer systems17;18;19. The prepara-\ntion of alloy like and composition stable nanojunc-\ntions composed of CoandGdbetween cobalt leads\nis hence possible in principle experimentally due to\nthe method of controlled interdiffusion process.\nIt should be noted that there has been at-\ntempts in the past to model the magnetic proper-\nties of bulk and layered cobalt-gadolinium systems\n13;20;21usingthe MFT method. These model cal-\nculations have been performed by adjusting in gen-\neral the MFT results to fit the experimental data,\nusing the cobalt spin SCoas a fitting parameter. In\nsome of the calculations where SCois assigned its\nfundamental value, the overall fit function for the\nmagnetization with temperature does not give bet-\nter agreement with the experimental data. Further-\nmore, theasymmetricchoice of nearestneighbor ex-\nchange constants for cobalt JCo−Coand gadolinium\nJGd−Gdis made in these references to reduce the\nnumber of adjustable parameters, but without giv-\ning any fundamental justification. To add to this\ncomplex situation, there is a wide array of exper-\nimental values of exchange constant for the cobalt\nJCo−Coand gadolinium JGd−Gdare available in the\nliterature from different types of measurements22,\nwhich does not help to clarify the situation for ad-\nvanced modeling.\nIn our previous work23, we computed the\nballistic and scattering transport properties of\nspin waves (SW) incident from cobalt leads, on\nto the embedded ultrathin ferrimagnetic cobalt-\ngadolinium [ Co1−cGdc]ℓnanojunction systems be-\ntween the leads. The nanojunction [ Co1−cGdc]ℓis\nprinciple a randomly disordered alloy with varied\nhcp atomic palnes ℓbetween matching hcp planes\nof theColeads, at known stable concentrations\nc≤0.5 for this nano-alloy system. To be able to\ncarry out these computations it was necessary to\ncompute the sublattice magnetizations and mag-\nnetic exchange constants in this system24.\nIn the present work we have modeled the\nsublattice magnetizations and magnetic exchange\nconstants of the alloy layered nanostructures\n[Co1−cGdc]ℓ′[Co]ℓ[Co1−cGdc]ℓ′sandwiched between\nsemi-infinite cobalt leads, at concentrations c≤\n0.5. These triple-nanostructure systems are more\ncomplex than the previously studied single-\nnanostructure systems24. The present work is\nhence motivated by the objective to present a\nmore complex computational modeling of the\nsalient nano-magnetic properties of the triple-Sublattice magnetizations of nano-magnetic layered mater ials3\nnanostructures for fundamental interest, and the\nballistic transport and scattering of spin waves in-\ncident from cobalt leads on complex ferrimagnetic\ncobalt-gadolinium nanojunction systems25. Such\ncomplex systems have as it turns out a richer and\nwider range of spin wave filtering properties. The\ncomplex embedded triple-nanostructures are de-\nnoted symbolically henceforth by [ ℓ′ℓℓ′] for conve-\nnience, corresponding to the alternating alloy and\npure nanostructures. The basal hcp (0001) atomic\nplanesofthe ...Co][ℓ′ℓℓ′][Co...nanojunctionsystems\nare normal to the direction of the c-axis itself con-\nsidered to be along the direction of the leads.\nThe alloy layered [ ℓ′ℓℓ′] nanostructures under\nconsideration are ultrathin ∼1.5 nm composite\ncobalt-gadolinium alloy systems sandwiched be-\ntween Co leads, and are hence different from bulk\nalloy and multilayer systems8;21;13. Also, due to\nthe absence of first principle calculations for Co-\nCo and Co-Gd exchange in the alloy layered [ ℓ′ℓℓ′]\nnanostructures.There is hence effectively a need for\nreliable data for the exchange and sublattice mag-\nnetizations in these systems to be able to develop\nmodelingstudiesfor the ...Co][ℓ′ℓℓ′][Co...nanojunc-\ntions which are key elements for ballistic spin wave\ntransport in magnonic devices23;26;27;28. This\nneed has motivated our EFT calculations to deter-\nmine such data with no fitting parameters, using\nbasic values SCo= 1 and SGd= 7/2 as the spin\nreferences at absolute 0 K.\nThe structure of the paper is as follows. In sec-\ntion 2, the EFT Ising spin method with experimen-\ntal data is computed for the reliable JCo−Coand\nJGd−Gdexchange for the purecrystalline cobalt and\ngadolinium materials. These are then attributed to\nnearest neighbor Co−CoandGd−Gdinterac-\ntions in the ...Co][ℓ′ℓℓ′][Co...nanojunction systems\nfor eutectic stable concentrations c≤0.5. The com-\nbined EFT and MFT methods are presented in sec-\ntion 3, to compute sublattice magnetizations for the\ncobalt and gadolinium sites on the individual hcp\nbasal atomic planes of the alloy layered [ ℓ′ℓℓ′] lamel-\nlar nanostructures as a function of temperature,\nwith thicknesses [2′22′] and [3′33′], and for differ-\nent alloy concentrations c. The sublattice magne-\ntizations and corresponding ferrimagnetic compen-\nsation temperatures are shown in this section. The\noverall discussions and conclusions are presented in\nsection 4.2. EFT modeling of pure CoandGd\nsystems\nTheEFT modeling incorporates the contribution of\nthe single site spin correlations to the order param-\neter, and hence known to be superior to the MFT.\nWe use it in the present work to model and calcu-\nlate the exchange constant for CoandGdcrystals\nover their ordered nano-magnetic phase, by com-\nparing the EFT magnetization results and Curie\ntemperatures with the experimental data29. The\nexchange for CoandGdcrystals are then calcu-\nlated by using our EFT constitutive relations30\nkTc/zJS(S+ 1) = 0 .3127 and kTc/zJS(S+ 1) =\n0.3162, valid for cobalt and gadolinium, respec-\ntively. The EFT exchange JCo−CoandJGd−Gdare\ningoodagreementwiththemeanvaluesofexchange\nconstant obtainedfromextensiveexperimental data\n22forCoandGd. The calculated EFT magne-\ntization and exchange results for cobalt are then\nused for the calculations of sublattice magnetiza-\ntions of the alloy layered ...Co][ℓ′ℓℓ′][Co...nanos-\ntructure between Co leads, by using the MFT\nmethod20;21;13. The schematic representation\nfor the consecutive planes of cobalt-gadolinium\nVCA alloy and pure cobalt plane of nanojunction\n[Co1−cGdc]2[Co]2[Co1−cGdc]2between crystalline\ncobalt leads are shown in Fig.1.\nThe Ising spin Hamiltonian Hin the absence\nof local spin anisotropy and Zeeman effects may be\nexpressed as,\nH=−J/summationdisplay\nSiz.Sjz, (1)\nwhere,/angb∇acketlefti,j/angb∇acket∇ightrepresents sum over nearest neighbors\nin the crystal, and Jis the nearest neighbors mag-\nnetic exchange constant that induces spin order\nalong a selected z-axis. The coordination number\nofCoandGdisz= 12, and present negligible\nanisotropy in the bulk material as compared to\ntheexchange. TheHamiltonian Eq.(1)forcomputa-\ntional purposes may be written in more useful form\nas,\nH=/summationdisplay\ni/summationdisplay\nj(−JSjz)Siz≡/summationdisplay\niHi(x)≡/summationdisplay\ni−x Siz,\n(2)\nThe thermodynamic canonical averages for any\ndesiredspin operator “Op” may be calculated by\nthe effective filed theory method using the Van der\nWearden’s (VdW) operator exp( JSz∇). TheMath-\nematica code formulation for any given characteris-4Vinod Ashokan; A. Khater and M. Abou Ghantous\nFig. 1. Schematic representation for the consecutive plane s of cobalt-gadolinium VCA alloy and pure cobalt plane of nan o-\njunction [ Co1−cGdc]2[Co]2[Co1−cGdc]2between crystalline cobalt leads. The hcp crystal c-axis is normal to the symmetry\n(0001) atomic planes.\ntic function fOp(x) of the spin system may be ex-\npressed as\nfOp(x) =Tr(Op.MatrixExp[ −Hi(x)/kT])\nTr(MatrixExp[ −Hi(x)/kT]),(3)\nwhere the Van der Wearden’s (VdW) operator for\ncobalt with S= 1 is given as31,\nexp(JSz∇) =S2\nzcosh(J∇)+Szsinh(J∇)+1−S2\nz.\n(4)\nThe differential operator ∇=∂/∂xoperate with\nproperty fOp(x)|x→0=fOp(x+\na)|x→0=fop(a). The EFT method calcula-\ntions are discussed in detail in the earlier works\n31;32;33;34. With the help of Matrix quantum\nmechanics all required averages can be evaluated\nwith symbolic and numerical procedures.\nThe thermodynamic canonical averages are\nrepresented by =<(exp(JSz∇))z>\nfOp(x)|x→0. The canonical averages are confined to\nsingle-sitespinvariables σ=< Sz>andq=< S2\nz>\nwith reference to their basic spin values S= 1 for\nCoandS= 7/2 forGdatT= 0K. In the com-\nputational procedure desired decoupling approxi-\nmation is used, which is equivalent to neglecting\nthe site-site correlations < SizSjz>but preserv-\ning the single-site correlations < SizSiz>. Further-\nmore, this decoupling approximation makes EFT\napproachquiteeffectivetocomputethesalient mag-\nnetic properties of the systems.The comparison between the normalized mag-\nnetizations for the CoandGdcrystals calculated\nusing EFT method, and by using the EFT consti-\ntutive relations30,kTc/zJS(S+ 1) = 0 .3127 for\nCo(S= 1,z= 12), and kTc/zJS(S+1) = 0 .3162\nforGd(S= 7/2,z= 12), with their experimental\nmeasurements29;35;36;37, yield their respective\nCurie temperatures, that is 119.4 meV and 25.2\nmeV24. It is striking to note that for Cowith\nspinS= 1, the EFT calculated magnetization and\nCurie temperature agree with the experimentally\nobserved mean value of exchange constant JCo−Co\n= 15.9 meV22. Similarly, the hcp Gdwith the\nspinS= 7/2, the EFT calculated magnetization\nand Curie temperature agree with the correspond-\ning experimentally measured exchange JGd−Gd=\n0.42meV.29.TheEFTcalculated exchange JCo−Co\nand spin variable σCofor sites on the cobalt leads\nseed from the interfaces inwards, to calculate the\nsublattice magnetization using MFT of embedded\nalloy layered [ ℓ′ℓℓ′] nanostructure.\n3. Sublattice magnetizations of the\nalloy layered [ℓ′ℓℓ′]ferrimagnetic\nnanostructures between cobalt\nleads\nInthepresentcomputationalmodelingthealloyhcp\natomic planes of the layered [ ℓ′ℓℓ′] nanostructure\nsandwiched between cobalt leads, are modeled asSublattice magnetizations of nano-magnetic layered mater ials5\ncrystalline atomic planes. On their hcp lattice there\nis a random homogeneous distributions of Coand\nGdatoms. Any random site is considered to have\nthe usual six nearest neighbors in its hcp (0001)\nbasal plane, and another six neighbors on the two\nadjacent planes. The system is made of alternat-\ning hcp (0001) atomic planes, and the structural\nmorphology of the two interfaces between the leads\nand the layered nanostructure are abrupt and crys-\ntalline. The advent of advanced experimental tech-\nniques for Co/Gdmultilayer systems,12;11, per-\nmits minimizing the interface roughness, and the\ncontrol of the atomic interdiffusion towards stable\neutectic compositions c= 0.1 to 0.5.\nTo calculate the sublattice magnetization for\nthe individual basal atomic hcp planes of the al-\nloy layered [ ℓ′ℓℓ′] nanostructures by MFT, the Bril-\nlouin’s functions are used to calculate initially\nthe different spin variables σ(n′)\nαfor then′th lay-\nered atomic plane. To simplify the notation we\nsystematically call σ(n′)\nαas the thermodynamic\ncanonical spin variable so that σ(n′)\nα=Sα.BS≡\nBα(Sα,T,H(n′)\nα), where Sαrepresents the funda-\nmental atomic spin and αnamely for the Co and\nGd atoms, and BSis the Brillouin function. The\nthermodynamic canonical spin variable is given as,\nσ(n′)\nα=\n2Sα+1\n2Coth/parenleftigg\n2Sα+1\n2SαH(n′)\nα\nkT/parenrightigg\n−1\n2Coth/parenleftigg\n1\n2SαH(n′)\nα\nkT/parenrightigg\n(5)\nwhereH(n′)\nαrepresents the molecular field energy\nfor the element αin the atomic plane n′due to its\ninteraction with its z= 12 nearest neighbors. The\nkTrepresentsthermalenergyandtheeffective mag-\nnetic moment per site is ¯M(n′)in then′th plane, is\ngiven as in units of Bohr magnetons by\n¯M(n′)/µB= (1−c)g(n′)\nCoσ(n′)\nCo+cg(n′)\nGdσ(n′)\nGd.(6)\ng(n′)\nαare the g factors for the alloy element on the\nn′th plane. The magnetization for the n′th atomic\nplane is calculated by multiplying ¯M(n′)by the\nnumber of sites per unit volume for the atomic\nplane.3.1.Alloy layered\n[Co1−cGdc]2[Co]2[Co1−cGdc]2\nferrimagnetic nanostructure\nbetween cobalt leads\nConsider in this subsection the embedded layered\n[2′22′] nanostructure between cobalt leads. For the\nCoandGdatomintheferrimagneticalloyed atomic\nplanes are found with the respective probabilities\n(1−c) andc. The molecular field energy usingMFT\nfor aCoatom on the 1st hcp basal plane of the lay-\nered nanostructure at the interface with the cobalt\nlead, may hence be expressed as\nH(1)\nCo= (3σ(B)\nCoJcc)+6[(1−c)σ(1)\nCoJcc+cσ(1)\nGdJcg]\n+3[(1−c)σ(2)\nCoJcc+cσ(2)\nGdJcg]. (7)\nThe exchange interactions are denoted by the\nsimplified notation JCoCo≡Jcc,JGdGd≡Jgg, and\nJCoGd≡Jcg. Equally, the molecular field energy for\naGdatom on the 1st hcp basal plane of the al-\nloy layered nanostructure at the interface with the\ncobalt lead, is\n(a)\nCo\nGdc=0.1 c=0.5\nc=0.5 c=0.1\nstep 0.1\nmeV2|22|\n020406080100120-3-2-101\nkTσCo1σGd16Vinod Ashokan; A. Khater and M. Abou Ghantous\nCo\nGd(b)\nc=0.1\nc=0.5c=0.5\nc=0.1\nmeVstep 0.1\n2|22|\n020406080100120-3-2-101\nkTσCo2σGd2\nFig. 2. Calculated spin variables σCoandσGd, forCoand\nGdsites on the 1st (a) and 2nd (b) hcp basal (0001) planes,\nof the alloy layered [ Co1−cGdc]2[Co]2[Co1−cGdc]2ferrimag-\nnetic nanostructure between cobalt leads, for different all oy\nconcentrations c, as a function of kTin meV. The down (up)\narrows in each figure correspond to the trend of the σspin\nvariations for the Co(Gd) sites with cstep changes.\nH(1)\nGd= (3σ(B)\nCoJcg)+6[(1−c)σ(1)\nCoJcg+cσ(1)\nGdJgg]\n+3[(1−c)σ(2)\nCoJcg+cσ(2)\nGdJgg]. (8)\nIn the present formulation, the seeding spin\nvalue for the lead Coatom at the interface with the\nalloy layered nanostructure is represented by σ(B)\nCo,\nwhich is obtained singularly from the EFT calcula-\ntions described in detail in section 2.\nIn contrast, the molecular field for a Coatom\non the 2nd hcp basal plane of the layered nanos-\ntructure inwards from the 1st, is\nH(2)\nCo= 3[(1−c)σ(1)\nCoJcc+cσ(1)\nGdJcg]+6[(1−c)σ(2)\nCoJcc\n+cσ(2)\nGdJcg]+3σ(3)\nCoJcc (9)\nSimilarly, the corresponding molecular field for a\nGdatom on the 2nd hcp basal plane of the layered\nnanostructure, is\nH(2)\nGd= 3[(1−c)σ(1)\nCoJcg+cσ(1)\nGdJgg]+6[(1−c)σ(2)\nCoJcg\n+cσ(2)\nGdJgg]+3σ(3)\nCoJcg (10)M1/\u0001Bc=0.1\nc=0.5\nM2/\u0000B\n020406080100120-1.0-0.50.00.5\nkTM1,2(persite) / μB\nFig. 3. Calculated magnetic moments per site for sites on\nthe 1st (solid curves) and 2nd (dotted curves) hcp basal\n(0001) planes ofthe alloy layered ferrimagnetic nanostruc ture\n[Co1−cGdc]2[Co]2[Co1−cGdc]2between cobalt leads. They\npresent small differences only at the high temperature kT\nend of the ordered phase. The down arrows follow the varia-\ntion trend for the magnetic moments per site with the cstep\nchanges, on the 1st (solid arrow) and 2nd (dotted arrow) hcp\nbasal (0001) planes.\nThehcp basal planes of the pure[ Co]2layer be-\ntween the alloy layers [ Co1−cGdc]2, are designated\nrespectively as the 3rd and 4th atomic planes. Us-\ning the symmetry properties of the layered [2′22′]\nnanostructure, we note that σ(3)\nCo≡σ(4)\nCo. The corre-\nsponding molecular field for a Coatom on the 3rd\nhcp basal plane is hence\nH(3)\nCo= 3[(1−c)σ(2)\nCoJcc+cσ(2)\nGdJcg]+9σ(3)\nCoJcc(11)\nThe above equations can be put into matrix\nform\n\nH(1)\nCo\nH(1)\nGd\nH(2)\nCo\nH(2)\nGd\nH(3)\nCo\n=\nA1\nA2\n0\n0\n0\n+\nx1x2x3x4x5\ny1y2y3y4y5\nu1u2u3u4u5\nv1v2v3v4v5\nz1z2z3z4z5\n\nσ(1)\nCo\nσ(1)\nGd\nσ(2)\nCo\nσ(2)\nGd\nσ(3)\nCo\n\n(12)Sublattice magnetizations of nano-magnetic layered mater ials7\nand the coefficients matrix is identical to\nx1x2x3x4x5\ny1y2y3y4y5\nu1u2u3u4u5\nv1v2v3v4v5\nz1z2z3z4z5\n\n≡\n2βJcc6cJcgβJcc3cJcg0\n2βJcg6cJggβJcg3cJgg0\nβJcc3cJcg2βJcc6cJcg3Jcc\nβJcg3cJgg2βJcg6cJgg3Jcg\n0 0 βJcc3cJcg9Jcc\n(13)\nwhereA1= 3σ(B)\nCoJcc,A2= 3σ(B)\nCoJcgandβ=\n3(1−c). Using Eqs.(7) to (11), and the spin vari-\nablesσ(n′)\nαformat given by Eq.(5), it follows that\nEq.(12) represents a nonlinear equations and to be\nsolved for the spin variables\nM1/μB\nM2/μB\nM3/μB meVc=0.1\nc=0.5\n3|33|\n020406080100120-1.0-0.50.00.5\nkTM1,2,3(persite) / μB\nFig. 4. Calculated magnetic moments per site for the 1st,\n2nd, and 3rd hcp basal atomic planes of the alloy layered\n[Co1−cGdc]3[Co]3[Co1−cGdc]3ferrimagnetic nanostructure.\nThe magnetic moments per site on the 1st and 3rd hcp planes\n(discontinuous and continues curves, respectively) are qu ite\nsimilar throughout the temperature range of the ordered\nphase. They differ significantly from the magnetic moments\non the 2nd hcp plane (continues curves). The down arrows\ncorrespond to the trend of the magnetic moment variations\nwithcstep changes.\n\nσ(1)\nCo\nσ(1)\nGd\nσ(2)\nCo\nσ(2)\nGd\nσ(3)\nCo\n=\nBCo(SCo,T,H(1)\nCo)\nBGd(SGd,T,H(1)\nGd)\nBCo(SCo,T,H(2)\nCo)\nBGd(SGd,T,H(2)\nGd)\nBCo(SCo,T,H(3)\nCo)\n(14)Solving the above equations numerically, we\ncalculate the spin variables σ(1)\nCo,σ(1)\nGd,σ(2)\nCo,σ(2)\nGdand\nσ(3)\nCoas a function of temperature, for any given al-\nloy concentration c. Note that the symmetry of the\nsystem imposes the following equalities for the spin\nvariables\n/parenleftigg\nσ(1)\nCo\nσ(1)\nGd/parenrightigg\n≡/parenleftigg\nσ(6)\nCo\nσ(6)\nGd/parenrightigg\n,/parenleftigg\nσ(2)\nCo\nσ(2)\nGd/parenrightigg\n≡/parenleftigg\nσ(5)\nCo\nσ(5)\nGd/parenrightigg\n,\n/parenleftig\nσ(3)\nCo/parenrightig\n≡/parenleftig\nσ(4)\nCo/parenrightig\n(15)\nThe calculated spin variables σ(n′)\nCo,σ(n′)\nGdfor the\nnominal n′= 1 andn′= 2 hcp basal planes are pre-\nsented inFig.2, as afunctionof temperatureandfor\neutectic concentrations c= [0.1,0.5] in steps of 0.1.\nFurther, Eq.(6), and g(n′)\nCo≡gCo= 2.2,g(n′)\nGd≡\ngGd= 2 for all n′, yield the magnetic moments per\nsite on the 1st and 2nd hcp basal planes as a func-\ntion of temperature. These are presented for com-\nparison in Fig.3, where a small interesting differ-\nence is observed at the high temperature kTend\nof the ordered phase. Compensation temperatures\nkTcomp<21 meV, are observed for the ferrimag-\nnetic hcp planes for eutectic stable concentrations\nin the range 0 .23< c <0.5.\n3.2.Alloy layered\n[Co1−cGdc]3[Co]3[Co1−cGdc]3\nferrimagnetic nanostructure\nbetween cobalt leads\nThe layered [3′33′] nanostructures under considera-\ntion are symmetric about the origin here taken as\nthe hcp plane n= 0. In this system the symmetry\nproperties to be used are\n/parenleftigg\nσ(1)\nCo\nσ(1)\nGd/parenrightigg\n≡/parenleftigg\nσ(9)\nCo\nσ(9)\nGd/parenrightigg\n,/parenleftigg\nσ(2)\nCo\nσ(2)\nGd/parenrightigg\n≡/parenleftigg\nσ(8)\nCo\nσ(8)\nGd/parenrightigg\n,\n/parenleftigg\nσ(3)\nCo\nσ(3)\nGd/parenrightigg\n≡/parenleftigg\nσ(7)\nCo\nσ(7)\nGd/parenrightigg\n,and/parenleftig\nσ(4)\nCo/parenrightig\n≡/parenleftig\nσ(6)\nCo/parenrightig\n(16)\nSimilarly, as in the previous case [2′22′], the\nequivalent molecular field energy results can be cast\nhere in matrix form as8Vinod Ashokan; A. Khater and M. Abou Ghantous\n\nH(1)\nCo\nH(1)\nGd\nH(2)\nCo\nH(2)\nGd\nH(3)\nCo\nH(3)\nGd\nH(4)\nCo\nH(5)\nCo\n=\nA1\nA2\n0\n0\n0\n0\n0\n0\n\n+\n2βJcc6cJcgβJcc3cJcg0 0 0 0\n2βJcg6cJggβJcg3cJgg0 0 0 0\nβJcc3cJcg2βJcc6cJcgβJcc3cJcg0 0\nβJcg3cJgg2βJcg6cJggβJcg3cJgg0 0\n0 0 βJcc3cJcg2βJcc6cJcg3Jcc0\n0 0 βJcg3cJgg2βJcg6cJgg3Jcg0\n0 0 0 0 βJcc3cJcg6Jcc3Jcc\n0 0 0 0 0 0 6 Jcc6Jcc\n\n×\nσ(1)\nCo\nσ(1)\nGd\nσ(2)\nCo\nσ(2)\nGd\nσ(3)\nCo\nσ(3)\nGd\nσ(4)\nCo\nσ(5)\nCo\n(17)\nwhereA1(kT) = 3MCo(kT)JCo−CoandA2(kT) =\n3MCo(kT)JCo−Gd. This yields the new irreducible\nvariables\n\nσ(1)\nCo\nσ(1)\nGd\nσ(2)\nCo\nσ(2)\nGd\nσ(3)\nCo\nσ(3)\nGd\nσ(4)\nCo\nσ(5)\nCo\n=\nBCo(SCo,T,H(1)\nCo)\nBGd(SGd,T,H(1)\nGd)\nBCo(SCo,T,H(2)\nCo)\nBGd(SGd,T,H(2)\nGd)\nBCo(SCo,T,H(3)\nCo)\nBGd(SGd,T,H(3)\nGd)\nBCo(SCo,T,H(4)\nCo)\nBCo(SCo,T,H(5)\nCo)\n(18)c=0.47\n2|22|3|33|\nmeV(a)\n020406080100120-2.0-1.5-1.0-0.50.00.5\nkTM1(persite) / μB\n3|33|\n2|22|c=0.47 (b)\nmeV\n020406080100120-2.0-1.5-1.0-0.50.00.5\nkTM2(persite) / μB\nFig. 5. Calculated effective magnetic moments per site for\nthe nominal: (a) 1st alloyed, n′= 1, and (b) 2nd alloyed,\nn′= 2, hcp basal planes,\nfor the alloy layered [ Co0.53Gd0.47]2[Co]2[Co0.53Gd0.47]2\nand [Co0.53Gd0.47]3[Co]3[Co0.53Gd0.47]3nanostructures be-\ntween cobalt leads.\nThe above nonlinear equations can be solve nu-\nmerically to obtain the spin variables and magnetic\nmomentspersite,ontheindividualhcpbasalplanes\nof the alloy layered magnetic [3′33′] nanostructure.\nFig.4 presents the calculated results for the mag-\nnetic moments on the 1st, 2nd, and 3rd hcp atomic\nplanes for this system. The magnetic moments per\nsite for the 1st and 3rd hcp planes (discontinuous\nand continues curves, respectively) are quite similar\nthroughout the temperature range of the ordered\nphase. Together, they differ significantly from theSublattice magnetizations of nano-magnetic layered mater ials9\nmagnetic moments per site for the 2nd hcp plane,\nthroughout the temperature range of the ordered\nphase.\nAs defined, the fundamental atomic spins are\nS= 1 forCo, andS= 7/2 forGd, where we con-\nsider the spins to be up for Coand down for Gd,\nin the ferrimagnetic alloy. Note that the Gdfun-\ndamental spin is 3.5 times that of Co. However,\nthe magnetic exchange Jggbetween GdandGdis\nweaker than Jccbetween CoandCo. By studying\nequations7,8,9,and10, onecanseethatthemolec-\nular fields for CoandGdsites vary with tempera-\ntureandcanchangesigns.Thecompetitionbetween\nthese molecular fields along the entire temperature\nrange determines the magnetizations per site, M;\nin the low temperature regime they increase then\nreach a maximum value before losing their magne-\ntized phase. See figures 3 and 4.\nThe computational model is general and can be\nextended to treat individual hcp atomic planes of\nalloy layered [ Co1−cGdc]p[Co]q[Co1−cGdc]rnanos-\ntructures, with r,p,q≥1. This procedure may\nbe generalized to larger layered nanostructures be-\ntween semi-infinite Coleads. It has been observed\nthat while increasing r,p,qthe results for the sub-\nlattice magnetization properties in the core atomic\nplanes tend to limiting solutions.\nFig. 6. Calculated spin variables σ(n′)\nCoand their detailed\nvariations as a function of temperature on the nominal n′\ncobalt hcp basal planes inside the alloy layered ferrimag-\nnetic nanostructures [ Co0.53Gd0.47]2[Co]2[Co0.53Gd0.47]2\nand[Co0.53Gd0.47]3[Co]3[Co0.53Gd0.47]3incomparison with\nthe temperature variation of the spin variable σ(B)\nCoon the\ncobalt leads; see details in the text.3.3.Alloy layered\n[Co0.53Gd0.47]ℓ′[Co]ℓ[Co0.53Gd0.47]ℓ′\nferrimagnetic nanostructure\nbetween cobalt leads\nThis alloy layered magnetic nanostructure between\ncobalt leads, at the characteristic eutectic concen-\ntrationc= 0.47, is particularly interesting since\nCo/Gdmagnetic multilayers at the same composi-\ntion have been reported to be very stable,12. We\nhave appliedhencethe EFT- MFT model approach\nto deduce the spin variables σ(n′)\nCoandσ(n′)\nGd, and\nthe effective magnetic moments per site, for the in-\ndividual hcp basal planes of layered ferrimagnetic\nnanostructures between cobalt leads, as a function\nof temperature, eutectic concentration, and thick-\nnessesℓ= 2 and 3. The integer n′numbers the hcp\nplanes from 1 to 6 for ℓ= 2, and from 1 to 9 for\nℓ= 3.\nThe effective magnetic moments per site calcu-\nlated in the units of Bohr magnetons for the alloyed\nnominal 1st, n′= 1, and 2nd, n′= 2, hcp basal\nplanes for the alloy layered ferrimagnetic nanos-\ntructures [ Co0.53Gd0.47]2[Co]2[Co0.53Gd0.47]2and\n[Co0.53Gd0.47]3[Co]3[Co0.53Gd0.47]3arepresentedin\nFig.5. It is observed that the computed effective\nmagnetic moments per site as a function of tem-\nperature on the nominal n′= 1 hcp basal planes\ndo not vary significantly with increased thickness of\nnanostructure,see Fig.5(a). Thiscan beunderstood\nclearly since the corresponding matrix elements in\nEq.(17) do not change significantly with increasing\nthickness. In contrast, it is observed that the ef-\nfective magnetic moments per site on the nominal\nn′= 2 hcp basal planes do vary significantly with\nincreasing thickness, see Fig.5(b). This is expected\nphysically owing to the changes of the correspond-\ning effective molecular fields for CoandGdsites.\nThe observed variations start at ≈15 meV and per-\nsist for higher temperatures, including room tem-\nperature ≈26 meV.\nIt is also interesting to compute spin vari-\nablesσ(n′)\nCo, and their detailed variations as a func-\ntion of temperature on the nominal n′cobalt\nhcp basal planes inside the alloy layered ferri-\nmagnetic nanostructures, in comparison with the\nspin variable σ(B)\nCoon the cobalt leads as a\nfunction of temperature. This is done for the\n[Co0.53Gd0.47]ℓ′[Co]ℓ[Co0.53Gd0.47]ℓ′layered nanos-\ntructure between cobalt leads, for thicknesses ℓ= 2\nand 3. The calculated results are presented in Fig.6.10Vinod Ashokan; A. Khater and M. Abou Ghantous\nTable 1. Spin variable values σGd=< SGd>for theGdsites on atomic planes 1, 2, .., ℓ, of the\n2′22′and 3′33′nanojunction systems between cobalt leads, are given for st able eutectic compositions\nc≤0.5, at room temperature T=300K, using the theoretical EFT- MFT combined method. The spin\nvariable σCo=< SCo>is≈1 at room temperature for the Cosites throughout the system.\nConcentrations Spin variables σGd\n2′22′3′33′\nc [Co1−cGdc]2,L[Co1−cGdc]2,R[Co1−cGdc]3,L [Co1−cGdc]3,R\n0.1 -1.10 -1.10 -1.10 -1.10 -1.10 -1.08 -1.10 -1.10 -1.08 -1. 10\n0.2 -1.03 -1.03 -1.03 -1.03 -1.03 -1.00 -1.03 -1.03 -1.00 -1. 03\n0.3 -0.97 -0.97 -0.97 -0.97 -0.97 -0.91 -0.97 -0.97 -0.91 -0. 97\n0.4 -0.91 -0.91 -0.91 -0.91 -0.90 -0.81 -0.90 -0.90 -0.81 -0. 90\n0.5 -0.83 -0.83 -0.83 -0.83 -0.82 -0.71 -0.82 -0.82 -0.71 -0. 82\nAs is physically expected, the σ(B)\nCois≥σ(n′)\nCofor all\npureCohcp planes n′inside the layered thicknesses\nℓ, at all temperatures of the ordered ferrimagnetic\nphase. Also as expected, σ(5)\nCois≥σ(4)\nCofor the al-\nloy layered [ Co0.53Gd0.47]3[Co]3[Co0.53Gd0.47]3fer-\nrimagnetic nanostructure between cobalt leads, and\nboth are greater or equal to the σ(3)\nCoof the\n[Co0.53Gd0.47]2[Co]2[Co0.53Gd0.47]2layered nanos-\ntructure. Note that the n′= 3 pure cobalt plane\nfor the [2′22′] layered nanostructure is equivalent\nnominally to the n′= 4 pure cobalt plane for the\n[3′33′] layered nanostructure. The results confirm a\nphysical trend which is expected, and which would\nlead to limiting values with increasing thickness of\nthe layered ferrimagnetic structure between cobalt\nleads.\nWe emphasize that the basic physical variables,\nsuch as the exchange and sublattice magnetizations\nforCoandGdsitesfortheembeddedlayerednanos-\ntructures between cobalt leads, are necessary ele-\nments for the computations of the spin-dynamics\nof magnetic nanojunctions in the field of magnon-\nics, as for the ballistic magnon transport25. In Ta-\nble 1 we present an example for the calculated spin\nvariables < SCo>≈1 throughout the system, and\n< SGd>on theidentified atomic planes1, 2, .., ℓ,of\nthe layered ...Co][2′22′][Co...and...Co][3′33′][Co...\nnanostructures between cobalt leads, at room tem-\nperature T=300K and stable eutectic compositions\nc≤0.5, using the EFT-MFT combined method.\n4. Summary and conclusions\nIn this work, we model the salient sub-lattice mag-\nnetic properties of the alloy layered ferrimagnetic\n[Co1−cGdc]ℓ′[Co]ℓ[Co1−cGdc]ℓ′nanostructures be-\ntween magnetically ordered cobalt leads. In par-\nticular, sublattice magnetizations of the CoandGdsites on the individual hcp (0001) basal planes\nof the alloy layered lamellar nanostructures by\nusing EFT-MFT combined method. The effective\nmagnetic moments per site and sublattice mag-\nnetizations are plotted as a function of tempera-\nture and thicknesses of the lamellar nanostructure.\nThe computational model is general and can repre-\nsents other composite magnetic elements andlamel-\nlar nanostructures. The calculated magnetic ex-\nchange, and spin variables for gadolinium < SGd>\nand cobalt < SCo>, site on the identified hcp\n(0001) basal planes of the alloy layered ferrimag-\nnetic [Co1−cGdc]ℓ′[Co]ℓ[Co1−cGdc]ℓ′nanostructures\nbetween cobalt leads, are very important quanti-\nties for the self-consistent analysis of quantum spin\ndynamics system and the coherent magnon ballis-\ntic transport across such nanostructures. The cal-\nculated results are also important fo the applica-\ntions in the fields of magnonics. The Ising EFT\nmethod serves to determine the magnetic exchange\nconstants for CoandGdsites of purecrystals, char-\nacterized by their fundamental quantum spins, by\ncomparing EFT with the experimental data. By\nseeding the MFT results on the alloy lamellar ferri-\nmagnetic nanostructure by the EFT computations\nof the cobalt leads from the interface inwards, the\nsublattice magnetizations for the CoandGdsites\nin the nanostructure are computed.\nReferences\n1. A.A. Serga, A.V. Chumak, A. Andre, G.A. Melkov,\nA.N. Slavin, S.O. Demokritov, and B. Hillebrands,\nPhys. Rev. Lett. 99, (2007) 227202\n2. T. Schneider, A.A. Serga, B. Leven, B. Hillebrands,\nR.L. Stamps, and M.P. Kostylev, Appl. Phys. Lett.\n92, (2008) 022505\n3. V. V. Kruglyak, S. O. Demokritov and D. Grundler,\nJ. Phys. D: Appl. Phys. 43, (2010) 264001Sublattice magnetizations of nano-magnetic layered mater ials11\n4. K. Lee and S. Kim, J. Appl. Phys. 104, (2008)\n053909\n5. S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J.\nM. Daughton, S. von Moln´ ar, M. L. Roukes, A.\nY. Chtchelkanova and D. M. Treger , Science 294,\n(2004) 1488\n6. I. Zutic, Jaroslav Fabian and S. Das Sarma, Rev.\nMod. Phys. 76,(2004) 323\n7. L. Bogani and W. Wernsdofer, Nature Materials, 7,\n(2008) 179\n8. R.E. Camley and R. L. Stamps, J. Phys. Condens.\nMatter,5, (1993) 3727\n9. J.P. Andr´ es, J. L. Sacedo´ n, J. Colinoa) and J. M.\nRiveiro J. Appl. Phys. 87, (2000) 2483\n10. O.S. Anilturk and A.R. Koymen, Phys. Rev. B 68,\n(2003) 024430\n11. J.A. Gonzalez, J. Colino, J.P. Andres, M.A. Lopez\nde la Torre,J.M. Riveiro,PhysicaB 345, (2004)181\n12. J.P. Andr´ es, J. A. Gonzalez,T. P. A. Hase, B. K.\nTanner, and J. M. Riveiro, Phys. Rev. B 77, (2008)\n144407\n13. S. Demirtas, R. E. Camley, A. R. Koymena, Appl.\nPhy. Lett. 87, (2005) 202111\n14. Javier Hermosa-Mu¨ noz at. al Communications\nPhysics, 5, (2022) 26.\n15. P. Chaudhari, J. J. Cuomo, and R. J. Gambino,\nAppl. Phys. Lett. 22, (1973) 337\n16. P. Hansen, C. Clausen, G. Much, M. Rosenkranz,\nand K. Witter, J. App. Phys. 66, (1989) 756\n17. A. Khater, G. Le Gal, and T. Kaneyoshi, Phys. Let-\nters A171, (1992) 237\n18. M. Fresneau, G. Le Gal, and A. Khater, J. Mag.\nMag. Mat. 130, (1994) 63\n19. A. Khater, M. Abou Ghantous, and M. Fresneau, J.Mag. and Mag. Mat. 247, (2002) 305\n20. R.E. Camley and D.R. Tilley, Phys. Rev. B 37,\n(1988) 3413\n21. M. Mansuripur and M.F. Ruane, IEEE Trans.\nMagn.22, (1986) 33\n22. C.A.F. Vaz, J.A.C. Bland, and G. Lauhoff, Rep.\nProg. Phys. 71, (2008) 056501\n23. V. Ashokan, M. Abou Ghantous, D. Ghader, and A.\nKhater, J. Mag. Mag. Mat. 363, (2014) 66\n24. M. Abou Ghantous, A. Khater, V. Ashokan, D.\nGhader, J. Appl. Phys 113, (2013) 094303\n25. V. Ashokan, A. Khater, M. Abou Ghantous, D.\nGhader, J. Mag. Mag. Mat. 384, (2015) 18\n26. V. Ashokan, M. Abou Ghantous, D. Ghader, A.\nKhater, Thin Solid Films 616(2016) 6\n27. A. Khater, L. Saimb, R. Tigrineb, D. Ghader, Sur-\nface Science 672673 (2018) 47\n28. Farid Chelli, Boualem Bourahla and Antoine\nKhater, Int. J. of Mod. Phys. B, 34, (2020) 2050080\n29. M.D. Kuz’min, Phys. Rev. Lett. 94, (2005) 107204\n30. Elie A. Moujaes, A. Khater, M. Abou Ghantous, J.\nMag. and Mag. Mat. 391 (2015) 49\n31. J. W. Tucker,J. Phys.A: Math. Gen. 27, (1994)659\n32. A. Khater and M. Abou Ghantous, J. Mag. Mag.\nMat.323, (2011) 2717\n33. M. Abou Ghantous and A. Khater, J. Mag. Mag.\nMat.323, (2011) 2504\n34. R. Honmura and T. Kaneyoshi, J. Phys. C: Solid St.\nPhys.12, (1979) 3979\n35. H.P. Myers, and W. Sucksmith, Proc. R. Soc. Lon-\ndon A207, (1951) 427\n36. R. Pauthenet, J. Appl. Phys. 53, (1982) 8187\n37. H.E. Nigh, S. Legvold, and F.H. Spedding, Phys.\nRev.132, (1963) 1092." }, { "title": "1909.09085v1.Magnetization_dynamics_of_the_compensated_ferrimagnet__Mn__2_Ru__x_Ga_.pdf", "content": "Magnetisation dynamics of the compensated ferrimagnet Mn 2RuxGa\nG. Bon\fglio\nRadboud University, Institute for Molecules and Materials, 6525 AJ Nijmegen, The Netherlands\nK. Rode, K. Siewerska, J. Besbas, G. Y. P. Atcheson, P. Stamenov, and J.M.D. Coey\nCRANN, AMBER and School of Physics, Trinity College Dublin, Ireland\nA.V. Kimel, Th. Rasing, and A. Kirilyuk\nRadboud University, Institute for Molecules and Materials, 6525 AJ Nijmegen, The Netherlands and\nFELIX Laboratory, Radboud University, Toernooiveld 7c, 6525 ED Nijmegen, The Netherlands\nHere we study both static and time-resolved dynamic magnetic properties of the compensated\nferrimagnet Mn 2RuxGa from room temperature down to 10 K, thus crossing the magnetic compen-\nsation temperature TM. The behaviour is analysed with a model of a simple collinear ferrimagnet\nwith uniaxial anisotropy and site-speci\fc gyromagnetic ratios. We \fnd a maximum zero-applied-\n\feld resonance frequency of \u0018160 GHz and a low intrinsic Gilbert damping \u000b\u00180:02, making it a\nvery attractive candidate for various spintronic applications.\nI. INTRODUCTION\nAntiferromagnets (AFM) and compensated ferrimag-\nnets (FiM) have attracted a lot of attention over the last\ndecade due to their potential use in spin electronics1,2.\nDue to their lack of a net magnetic moment, they are\ninsensitive to external \felds and create no demagnetis-\ning \felds of their own. In addition, their spin dynamics\nreach much higher frequencies than those of their ferro-\nmagnetic (FM) counterparts due to the contribution of\nthe exchange energy in the magnetic free energy3.\nDespite these clear advantages, AFMs are scarcely\nused apart from uni-directional exchange biasing rela-\ntively in spin electronic applications. This is because\nthe lack of net moment also implies that there is no\ndirect way to manipulate their magnetic state. Fur-\nthermore, detecting their magnetic state is also compli-\ncated and is usually possible only by neutron di\u000braction\nmeasurements4, or through interaction with an adjacent\nFM layer5.\nCompensated, metallic FiMs provide an interesting al-\nternative as they combine the high-speed advantages of\nAFMs with those of FMs, namely, the ease to manipu-\nlate their magnetic state. Furthermore, it has been shown\nthat such materials are good candidates for the emerging\n\feld of All-Optical Switching (AOS) in which the mag-\nnetic state is solely controlled by a fast laser pulse6{8.\nA compensated, half-metallic ferrimagnet was \frst en-\nvisaged by van Leuken and de Groot9. In their model\ntwo magnetic ions in crystallographically di\u000berent po-\nsitions couple antiferromagnetically and perfectly com-\npensate each-other, but only one of the two contributes\nto the states at the Fermi energy responsible for elec-\ntronic transport. The \frst experimental realisation of\nthis, Mn 2RuxGa (MRG), was provided by Kurt et al.10.\nMRG crystallises in the XAHeusler structure, space\ngroupF\u001643m, with Mn on the 4 aand 4csites11.\nSubstrate-induced bi-axial strain imposes a slight tetrag-\nonal distortion, which leads to perpendicular magneticanisotropy. Due to the di\u000berent local environment of\nthe two sublattices, the temperature dependence of their\nmagnetic moments di\u000ber, and perfect compensation is\ntherefore obtained at a speci\fc temperature TMthat\ndepends on the Ru concentration xand the degree of\nbiaxial strain. It was previously shown that MRG ex-\nhibits properties usually associated with FMs: a large\nanomalous Hall angle12, that depends only on the mag-\nnetisation of the 4 cmagnetic sublattice13; tunnel magne-\ntoresistance (TMR) of 40 %, a signature of its high spin\npolarisation14, was observed in magnetic tunnel junc-\ntions (MTJs) based on MRG15; and a clear magneto-\noptical Kerr e\u000bect and domain structure, even in the ab-\nsence of a net moment16,17. Strong exchange bias of a\nCoFeB layer by exchange coupling with MRG through\na Hf spacer layer18, as well as single-layer spin-orbit\ntorque19,20showed that MRG combined the qualities of\nFMs and AFMs in spin electronic devices.\nThe spin dynamics in materials where two distinct\nsublattices are subject to di\u000bering internal \felds (ex-\nchange, anisotropy, . . . ) is much richer than that of a\nsimple FM, as previously demonstrated by the obers-\nvation of single-pulse all-optical switching in amorphous\nGdFeCo21,22and very recently in MRG8. Given that the\nmagnetisation of MRG is small, escpecially close to the\ncompensation point, and the related frequency is high,\nnormal ferromagnetic resonance (FMR) spectroscopy is\nunsuited to study their properties. Therefore, we used\nthe all-optical pump-probe technique to characterize the\nresonance frequencies at di\u000berent temperatures in vicin-\nity of the magnetic compensation point. This, together\nwith the simulation of FMR, make it possible to deter-\nmine the e\u000bective g-factors, the anisotropy constants and\ntheir evolution across the compensation point. We found,\nin particular, that our ferrimagnetic half-metallic Heusler\nalloy has resonance frequency up to 160 GHz at zero-\feld\nand a relatively low Gilbert damping.arXiv:1909.09085v1 [cond-mat.mtrl-sci] 19 Sep 20192\nFIG. 1. Net moment measured by magnetometry and coercive\n\feld measured by static Faraday e\u000bect. The upturn of the net\nmoment below T\u001850 K is due to paramagnetic impurities\nin the MgO substrate. TMis indicated by the vertical dotted\nline. As expected the maximum available applied \feld \u00160H=\n7 T is insu\u000ecient to switch the magnetisation close to TM.\nII. EXPERIMENTAL DETAILS\nThin \flm samples of MRG were grown in a `Sham-\nrock' sputter deposition cluster with a base pressure of\n2\u000210\u00008Torr on MgO (001) substrates. Further infor-\nmation on sample deposition can be found elsewhere23.\nThe substrates were kept at 250\u000eC, and a protective\n\u00183 nm layer of aluminium oxide was added at room tem-\nperature. Here we focus on a 53 nm thick sample with\nx= 0:55, leading to TM\u001980 K as determined by SQUID\nmagnetometry using a Quantum Design 5 T MPMS sys-\ntem (see FIG. 1). We are able to study the magneto-\noptical properties both above and below TM.\nThe magnetisation dynamics was investigated using an\nall-optical two-colour pump-probe scheme in a Faraday\ngeometry inside a \u00160Hmax= 7 T superconducting coil-\ncryostat assembly. Both pump and probe were produced\nby a Ti:sapphire femtosecond pulsed laser with a cen-\ntral wavelength of 800 nm, a pulse width of 40 fs and\na repetition rate of 1 kHz. After splitting the beam in\ntwo, the high-intensity one was doubled in frequency by\na BBO crystal (giving \u0015= 400 nm) and then used as\nthe pump while the lower intensity 800 nm beam acted\nas the probe pulse. The time delay between the two was\nadjusted by a mechanical delay stage. The pump was\nthen modulated by a synchronised mechanical chopper\nat 500 Hz to improve the signal to noise ratio by lock-in\ndetection. Both pump and probe beams were linearly\npolarized, and with spot sizes on the sample of 150 µm\nand 70 µm, respectively. The pump pulse hit the sample\nat an incidence angle of \u001910\u000e. After interaction with\nthe sample, we split the probe beam in two orthogonally\npolarized parts using a Wollaston prism and detect the\nchanges in transmission and rotation by calculating the\nFIG. 2. Comparison of hysteresis loops obtained by Faraday,\nAHE, and magnetometry recorded at room temperature. The\ntwo former were recorded with the applied \feld perpendicular\nto the sample surface, while for the latter we show results for\nboth \feld applied parallel and perpendicular to the sample.\nsum and the di\u000berence in intensity of the two signals.\nThe external \feld was applied at 75\u000eto the easy axis of\nmagnetization thus tilting the magnetisation away from\nthe axis. Upon interaction with the pump beam the mag-\nnetisation is momentarily drastically changed24and we\nmonitor its return to the initial con\fguration via remag-\nnetisation and then precession through the time depen-\ndent Faraday e\u000bect on the probe pulse.\nThe static magneto-optical properties were examined\nin the same cryostat/magnet assembly.\nIII. RESULTS & DISCUSSION\nA. Static magnetic properties\nWe \frst focus on the static magnetic properties as\nobserved by the Faraday e\u000bect, and compare them to\nwhat is inferred from magnetometry and the anomalous\nHall e\u000bect. In FIG. 2 we present magnetic hysteresis\nloops as recorded using the three techniques. Due to the\nhalf metallic nature of the sample, the magnetotrans-\nport properties depend only on the 4 csublattice. As the\nmain contribution to the MRG dielectric tensor in the\nvisible and near infrared arises from the Drude tail16,\nboth AHE and Faraday e\u000bect probe essentially the same\nproperties (mainly the spin polarised conduction band of\nMRG), hence we observe overlapping loops for the two\ntechniques. Magnetometry, on the other hand, measures\nthe net moment, or to be precise the small di\u000berence\nbetween two large sublattice moments. The 4 asublat-\ntice, which is insigni\fcant for AHE and Faraday here\ncontributes on equal footing. FIG. 2 shows a clear di\u000ber-\nence in shape between the magnetometry loop and the3\nFIG. 3. Time resolved Faraday e\u000bect recorded at T= 290 K\nin applied \felds ranging from 1 T to 7 T. After the initial\ndemagnetisation seen as a sharp increase in the signal at t\u0018\n0 ps, magnetisation is recovered and followed by precession\naround the e\u000bective \feld until fully damped. The lines are\n\fts to the data. The inset shows the experimental geometry\nfurther detailed in the main text.\nAHE or Faraday loops. We highlight here that the ap-\nparent `soft' contribution that shows switching close to\nzero applied \feld, is not a secondary magnetic phase, but\na signature of the small di\u000berences in the \feld-behaviour\nof the two sublattices. We also note that this behaviour\nis a result of the non-collinear magnetic order of MRG.\nA complete analysis of the dynamic properties therefore\nrequires knowledge of the anisotropy constants on both\nsublattices as well as the (at least) three intra and in-\nter sublattice exchange constants. Such an analysis is\nbeyond the scope of this article, and we limit our anal-\nysis to the simplest model of a single, e\u000bective uniaxial\nanisotropy constant Kuin the exchange approximation\nof the ferrimagnet.\nB. Dynamic properties\nWe now turn to the time-resolved Faraday e\u000bect and\nspin dynamics. Time-resolved Faraday e\u000bect data were\nrecorded at \fve di\u000berent temperatures 10 K, 50 K, 100 K,\n200 K and 290 K, with applied \felds ranging from 1 T to\n7 T.\nFIG. 3 shows the \feld-dependence of the Faraday ef-\nfect as a function of the delay between the pump and\nthe probe pulses, recorded at T= 290 K. Negative de-\nlay indicates the probe is hitting the sample before the\npump. After the initial demagnetisation, the magneti-\nsation recovers and starts precessing around the e\u000bec-\ntive \feld which is determined by the anisotropy and the\napplied \feld. The solid lines in FIG. 3 are \fts to the\ndata to extract the period and the damping of the pre-cession in each case. The \ftting model was an expo-\nnentially damped sinusoid with a phase o\u000bset. We note\nthat the apparent evolution of the amplitude and phase\nwith changing applied magnetic \feld is due to the quasi-\nresonance of the spectrum of the precessional motion\nwith the low-frequency components of the convolution\nbetween the envelope of the probe pulse and the phys-\nical relaxation of the system. The latter include both\nelectron-electron and electron-lattice e\u000bects. A rudimen-\ntary model based on a classical oscillator successfully re-\nproduces the main features of the amplitude and phase\nobserved.\nIn two-sublattice FiMs, the gyromagnetic ratios of the\ntwo sublattices are not necessarily the same. This is par-\nticularly obvious in rare-earth/transition metal alloys,\nand is also the case for MRG despite the two sublat-\ntices being chemically similar; they are both Mn. Due\nto the di\u000berent local environment however, the degree\nof charge transfer for the two di\u000bers. This leads to two\ncharacteristic temperatures, a \frst TMwhere the mag-\nnetic moments compensate, and a second TAwhere the\nangular momenta compensate. It can be shown that for\nthe ferromagnetic mode, the e\u000bective gyromagnetic ratio\n\re\u000bcan then be written25\n\re\u000b=M4c(T)\u0000M4a(T)\nM4c(T)=\r4c\u0000M4a(T)=\r4a(1)\nsubscripti= 4a;4cdenotes sublattice i,Mi(T)\nthe temperature-dependent magnetisation, and \rithe\nsublattice-speci\fc gyromagnetic ratio. \re\u000bis related to\nthe e\u000bective g-factor\nge\u000b=\re\u000bh\n\u0016B(2)\nwherehis the Planck constant and \u0016Bthe Bohr magne-\nton.\nThe frequency of the precession is determined by the\ne\u000bective \feld, which can be inferred from the derivative\nof the magnetic free energy density with respect to M.\nFor an external \feld applied at a given \fxed angle with\nrespect to the easy axis this leads to the Smit-Beljers\nformula26\n!FMR =\re\u000bvuut1\nM2ssin2\u001e\"\n\u000e2E\n\u000e\u00122\u000e2E\n\u000e\u001e2\u0000\u0012\u000e2E\n\u000e\u0012\u000e\u001e\u00132#\n(3)\nwhere\u0012and\u001eare the polar and azimuthal angles of the\nmagnetisation vector, and Ethe magnetic free energy\ndensity\nE=\u0000\u00160H\u0001M+Kusin2\u0012+\u00160M2\nscos2\u0012=2 (4)\nwhere the terms correspond to the Zeeman, anisotropy\nand demagnetising energies, respectively, and Msis the\nnet saturation magnetisation. It should be mentioned\nthat the magnetic anisotropy constant Kuis related to\nM, which is being considered constant in magnitude, via\nKu=\f\u00160M2\ns=2,\fa dimensionless parameter.4\nFIG. 4. Observed precession frequency as a function of the\napplied \feld for various temperatures. The solid lines are \fts\nto the data as described in the main text.\nBased on Eqs. (1) through (4) we \ft our entire data set\nwith\re\u000bandKuas the only free parameters. The exper-\nimental data and the associated \fts are shown as points\nand solid lines in FIG. 4. At all temperatures our simple\nmodel with one e\u000bective gyromagnetic ratio \re\u000band a\nsingle uniaxial anisotropy parameter Kureproduces the\nexperimental data reasonably well. The model systemat-\nically underestimates the resonance frequency for inter-\nmediate \felds, with the point of maximum disagreement\nincreasing with decreasing temperature. We speculate\nthis is due to the use of a simple uniaxial anisotropy in\nthe free energy (see Eq. 4), while the real situation is\nmore likely to be better represented as a sperimagnet. In\nparticular, the non-collinear nature of MRG that leads\nto a deviation from 180\u000eof the angle between the two\nsublattice magnetisations, depending on the applied \feld\nand temperature.\nFrom the \fts in FIG. 4 we infer the values of ge\u000band\nthe anisotropy \feld \u00160Ha=2Ku=Ms. The result is shown\nin FIG. 5. The anisotropy \feld is monotonically increas-\ning with decreasing temperature as the magnetisation\nof the 4csublattice increases in the same temperature\nrange. We highlight here the advantage of determining\nthis \feld through time-resolved magneto-optics as op-\nposed to static magnetometry and optics. Indeed the\nanisotropy \feld as seen by static methods is sensitive to\nthe combination of anisotropy and the netmagnetic mo-\nment, as illustrated in FIG. 1, where the coercive \feld\ndiverges as T!TM. In statics one would expect a di-\nvergence of the anisotropy \feld at the same temperature.\nThe time-resolved methods however distinguish between\nthe net and the sublattice moments, hence better re\rect-\ning the evolution of the intrinsic material properties of\nthe ferrimagnet.\nThe temperature dependence of the anisotropy con-\nstants was a matter for discussion for many years27,28.\nFIG. 5. E\u000bective g-factor,ge\u000b, and the anisotropy \feld\nas determined by time-resolved Faraday e\u000bect. ge\u000b, orange\nsquares, increases from near the free electron value of 2 to 4\njust belowTM, while the anisotropy \feld, blue triangles, in-\ncreases near-linearly with decreasing temperature. A M3\ft,\nred dashes line, of the anisotropy behaviour shows the almost-\nmetallic origin of it, indicating the dominant character of the\n4c sublattice.\nWritten in spherical harmonics the 3 danisotropy can\nbe expressed as, k2Y0\n2(\u0012) +k4Y0\n4(\u0012)29wherek2/\nM(T)3andk4/M(T)10. The experimental measured\nanisotropy is then, K2(T) =ak2(T)+bk4(T), withaand\nbthe contributions of the respective spherical harmonics.\nFIG. 5 shows that a reasonable \ft of our data is ob-\ntained with M(T)3which means, \frst, that the contri-\nbution of the 4thorder harmonic can be neglected, and\nsecond, that the contribution of the TMand 2ndsublat-\ntice is negligible, indicating the dominant character of\nthe 4c sublattice.\nIn addition, we should note here that the high fre-\nquency exchange mode was never observed on our exper-\niments. While far from TMits frequency might be too\nhigh to be observable, in the vicinity of TM, in contrast,\nits frequency is expected to be in the detection range.\nMoreover, given the di\u000berent electronic structure of the\ntwo sublattices, it is expected that the laser pulse should\nselectively excite the sublattice 4c, and therefore lead to\nthe e\u000bective excitation of the exchange mode. We argue\nthat it is the non-collinearity of the sublattices (see sec-\ntion III A) that smears out the coherent precession at\nhigh frequencies.\nThe e\u000bective gyromagnetic ratio, ge\u000b, shows a non-\nmonotonic behaviour. It increases with decreasing Tto-\nwardsTM, reaching a maximum at about 50 K before\ndecreasing again at T= 10 K. We alluded above to\nthe di\u000berence between the magnetic and the angular mo-\nmenta compensation temperatures. We expect that ge\u000b\nreaches a maximum when T=TA30, here between the\nmeasurement at T= 50 K and the magnetic compensa-\ntion temperature TM\u001980 K.5\nFIG. 6. Intrinsic and anisotropic broadening in MRG across\ntheTM. The inset shows the evaluation process of the two\ndamping parameters. A linear \ft is used to evaluate intercept\n(anisotropic broadening) and slope (intrinsic damping) of the\nfrequencies versus the inverse of the decay time. The data\npoint are obtained from the \ft of time-resolved Faraday e\u000bect\nmeasurements (an example is shown in Fig.4).\nFrom XMCD data11, we could estimate spin and or-\nbital moment components of the magnetic moments of\nthe two sublattices, what allowed us to derive the ef-\nfective g-factors for the sublattices as g4a= 2:05 and\ng4c= 2:00. In this case we expect the angular momentum\ncompensation temperature TAto be below TM, opposite\nto what is observed for GdFeCo21. Given this small dif-\nference however, TAandTMare expected to be rather\nclose to each other, consistent with the limited increase\nofge\u000bacross the compensation points.\nWe turn \fnally to the damping of the precessional mo-\ntion of Maround the e\u000bective \feld \u00160He\u000b. Damping is\nusually described via the dimensionless parameter \u000bin\nthe Landau-Lifshiz-Gilbert equation, and it is a measure\nof the dissipation of magnetic energy in the system. In\nthis model, \u000bis a scalar constant and the observed broad-\nening in the time domain is therefore a linear function of\nthe frequency of precession31{33. We infer \u000b0, the total\ndamping, from our \fts of the time-resolved Faraday e\u000bect\nas\u000b0= (\u001cd)\u00001, where\u001cdis the decay time of the \fts. We\nthen, for each temperature, plot \u000b0as a function of the\nobserved frequency and regress the data using a straight\nline \ft. The intrinsic \u000bis the slope of this line, while the\nintercept represents the anisotropic broadening.\nFIG. 6 shows the intrinsic damping \u000band the\nanisotropic broadening as a function of temperature.\nAnisotropic broadening is usually attributed to a vari-\nation of the anisotropy \feld in the region probed by the\nprobe pulse34. For MRG this is due to slight lateral vari-\nations in the Ru content xin the thin \flm sample. Such a\nvariation leads to a variation in e\u000bective TMandTAand\ncan therefore have a large in\ruence on the broadening asa function of temperature. Despite this, the anisotropic\nbroadening is reasonably low in the entire temperature\nrange above TM, and a more likely explanation for its\nrapid increase below TMis that the applied magnetic\n\feld is insu\u000ecient to completely remagnetize the sam-\nple between two pump pulses. As observed in Fig.5, the\nanisotropy \feld reaches almost 4 T at low temperature,\ncomparable to our maximum applied \feld of 7 T. The\nintrinsic damping \u000bis less than 0.02 far from TM, but\nincreases sharply at T= 100 K. We tentatively attribute\nthis to an increasing portion of the available power be-\ning transferred into the high-energy exchange mode, al-\nthough we underline that we have not seen any direct\nevidence of such a mode in any of the experimental data.\nIV. CONCLUSION\nWe have shown that the time-resolved Faraday e\u000bect\nis a powerful tool to determine the spin dynamic proper-\nties in compensated, metallic ferrimagnets. The high spin\npolarisation of MRG enables meaningful Faraday data to\nbe recorded even near TMwhere the net magnetisation\nis vanishingly small, and the dependence of the dynamics\non the sublattice as opposed to the net magnetic prop-\nerties provides a more physical understanding of the ma-\nterial. Furthermore, we \fnd that the ferromagnetic-like\nmode of MRG reaches resonance frequencies as high as\n160 GHz in zero applied \feld, together with a small in-\ntrinsic damping. This value is remarkable if compared\nto well-known materials such as GdFeCo which, at zero\n\feld, resonates at tens of GHz21or [Co/Pt] nmultilay-\ners at 80 GHz35but with higher damping. We should\nhowever stress that, in the presence of strong anisotropy\n\felds, higher frequencies can be reached. Example of that\ncan be found for ferromagnetic Fe/Pt with \u0019280 GHz\n(Ha= 10T)36, and for Heusler-like ferrimagnet (Mn 3Ge\nand Mn 3Ga) with\u0019500 GHz (Ha= 20T)37,38. Never-\ntheless, the examples cited above show a considerably\nhigher intrinsic damping compared to MRG. In addi-\ntion, it was recently shown that MRG exhibits unusu-\nally strong intrinsic spin-orbit torque20. Thus, taking\ninto account the material parameters we have determined\nhere, it seems likely it will be possible to convert a DC\ndriven current into a sustained ferromagnetic resonance\natf= 160 GHz, at least. These characteristics make\nMRG, as well as any future compensated half-metallic\nferrimagnet, particularly promising materials for both\nspintronics and all-optical switching.\nACKNOWLEDGMENTS\nThis project has received funding from the NWO pro-\ngramme Exciting Exchange, the European Union's Hori-\nzon 2020 research and innovation programme under grant\nagreement No 737038 `TRANSPIRE', and from Science6\nFoundation Ireland through contracts 12/RC/2278 AM-\nBER and 16/IA/4534 ZEMS.The authors would like to thank D. Betto for help ex-\ntractinghLiandhSi.\n1A. B. Shick, S. Khmelevskyi, O. N. Mryasov, J. Wunder-\nlich, and T. Jungwirth, Phys. Rev. B 81, 212409 (2010).\n2L. Caretta, M. Mann, F. B uttner, K. Ueda, B. Pfau,\nC. M. G unther, P. Hessing, A. Churikova, C. Klose,\nM. Schneider, D. Engel, C. Marcus, D. Bono, K. Bagschik,\nS. Eisebitt, and G. S. Beach, Nat. Nanotechnol. 13, 1154\n(2018).\n3E. V. Gomonay and V. M. Loktev, Low Temp. Phys. 40,\n17 (2014).\n4C. G. Shull and J. S. Smart, Phys. Rev. 76, 1256 (1949).\n5T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich,\nNat. Nanotechnol. 11, 231 (2016).\n6C. D. Stanciu, A. Tsukamoto, A. V. Kimel, F. Hansteen,\nA. Kirilyuk, A. Itoh, and T. Rasing, Phys. Rev. Lett. 99,\n217204 (2007).\n7S. Mangin, M. Gottwald, C.-H. Lambert, D. Steil, V. Uhl\u0013 \u0010,\nL. Pang, M. Hehn, S. Alebrand, M. Cinchetti, G. Mali-\nnowski, Y. Fainman, M. Aeschlimann, and E. E. Fullerton,\nNat. Mater. 13, 286 (2014).\n8C. Banerjee, N. Teichert, K. Siewierska, Z. Gercsi, G. Atch-\neson, P. Stamenov, K. Rode, J. M. D. Coey, and J. Besbas,\narXiv preprint arXiv:1909.05809 (2019).\n9H. van Leuken and R. A. de Groot, Phys. Rev. Lett. 74,\n1171 (1995).\n10H. Kurt, K. Rode, P. Stamenov, M. Venkatesan, Y. C.\nLau, E. Fonda, and J. M. D. Coey, Phys. Rev. Lett. 112,\n027201 (2014).\n11D. Betto, N. Thiyagarajah, Y.-C. Lau, C. Piamonteze, M.-\nA. Arrio, P. Stamenov, J. M. D. Coey, and K. Rode, Phys.\nRev. B 91, 094410 (2015).\n12N. Thiyagarajah, Y. C. Lau, D. Betto, K. Borisov, J. M.\nCoey, P. Stamenov, and K. Rode, Appl. Phys. Lett. 106,\n1 (2015).\n13C. Fowley, K. Rode, Y.-C. Lau, N. Thiyagarajah, D. Betto,\nK. Borisov, G. Atcheson, E. Kampert, Z. Wang, Y. Yuan,\nS. Zhou, J. Lindner, P. Stamenov, J. M. D. Coey, and\nA. M. Deac, Phys. Rev. B 98, 220406(R) (2018).\n14M.\u0014Zic, K. Rode, N. Thiyagarajah, Y.-C. Lau, D. Betto,\nJ. M. D. Coey, S. Sanvito, K. J. O'Shea, C. A. Fergu-\nson, D. A. MacLaren, and T. Archer, Phys. Rev. B 93,\n140202(R) (2016).\n15K. Borisov, D. Betto, Y. C. Lau, C. Fowley, A. Titova,\nN. Thiyagarajah, G. Atcheson, J. Lindner, A. M. Deac,\nJ. M. Coey, P. Stamenov, and K. Rode, Appl. Phys. Lett.\n108(2016), 10.1063/1.4948934.\n16K. Fleischer, N. Thiyagarajah, Y.-C. Lau, D. Betto,\nK. Borisov, C. C. Smith, I. V. Shvets, J. M. D. Coey, and\nK. Rode, Phys. Rev. B 98, 134445 (2018).\n17K. E. Siewierska, N. Teichert, R. Schfer, and J. M. D.\nCoey, IEEE Transactions on Magnetics 55, 1 (2019).18K. Borisov, G. Atcheson, G. D'Arcy, Y.-C. Lau, J. M. D.\nCoey, and K. Rode, Applied Physics Letters 111, 102403\n(2017), https://doi.org/10.1063/1.5001172.\n19R. E. Troncoso, K. Rode, P. Stamenov, J. M. D. Coey,\nand A. Brataas, Phys. Rev. B 99, 054433 (2019).\n20S. Lenne, Y.-C. Lau, A. Jha, G. P. Y. Atcheson, R. E.\nTroncoso, A. Brataas, J. Coey, P. Stamenov, and K. Rode,\narXiv preprint arXiv:1903.04432 (2019).\n21C. D. Stanciu, A. V. Kimel, F. Hansteen, A. Tsukamoto,\nA. Itoh, A. Kiriliyuk, and T. Rasing, Phys. Rev. B - Con-\ndens. Matter Mater. Phys. 73, 220402(R) (2006).\n22I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius,\nH. A. D urr, T. A. Ostler, J. Barker, R. F. Evans, R. W.\nChantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk, T. Rasing,\nand A. V. Kimel, Nature 472, 205 (2011).\n23D. Betto, K. Rode, N. Thiyagarajah, Y.-C. Lau,\nK. Borisov, G. Atcheson, M. ic, T. Archer, P. Stamenov,\nand J. M. D. Coey, AIP Advances 6, 055601 (2016),\nhttps://doi.org/10.1063/1.4943756.\n24B. Koopmans, J. J. M. Ruigrok, F. Dalla Longa, and\nW. J. M. de Jonge, Phys. Rev. Lett. 95, 267207 (2005).\n25R. K. Wangsness, Phys. Rev. 91, 1085 (153).\n26J. Smit and H. G. Beljers, R 263 Philips Res. Rep 10, 113\n(1955).\n27H. B. Callen and E. Callen, J. Phys. Chem. Solids 27, 1271\n(1966).\n28S. Vonsovskii, MAGNETISM. , vol. 2 (IPST, 1974).\n29M. Farle, Reports on Progress in Physics 61, 755 (1998).\n30A. Gurevich and G. Melkov, Magnetization Oscillations\nand Waves (Taylor & Francis, 1996).\n31G. Malinowski, K. C. Kuiper, R. Lavrijsen, H. J. M.\nSwagten, and B. Koopmans, Appl. Phys. Lett. 94, 102501\n(2009).\n32Y. Liu, L. R. Shelford, V. V. Kruglyak, R. J. Hicken,\nY. Sakuraba, M. Oogane, and Y. Ando, Phys. Rev. B\n81, 094402 (2010).\n33A. J. Schellekens, L. Deen, D. Wang, J. T. Kohlhepp,\nH. J. M. Swagten, and B. Koopmans, Appl. Phys. Lett.\n102, 082405 (2013).\n34J. Walowski, M. D. Kaufmann, B. Lenk, C. Hamann,\nJ. McCord, and M. M unzenberg, J. Phys. D. Appl. Phys.\n41, 164016 (2008), arXiv:0805.3495.\n35A. Barman, S. Wang, O. Hellwig, A. Berger, E. E. Fuller-\nton, and H. Schmidt, J. Appl. Phys. 101, 09D102 (2007).\n36J. Becker, O. Mosendz, D. Weller, A. Kirilyuk, J. C. Maan,\nP. C. M. Christianen, T. Rasing, and A. Kimel, Appl.\nPhys. Lett. 104, 152412 (2014).\n37S. Mizukami, A. Sugihara, S. Iihama, Y. Sasaki, K. Z.\nSuzuki, and T. Miyazaki, Applied Physics Letters 108,\n012404 (2016), https://doi.org/10.1063/1.4939447.\n38N. Awari, S. Kovalev, C. Fowley, K. Rode, R. Gallardo, Y.-\nC. Lau, D. Betto, N. Thiyagarajah, B. Green, O. Yildirim,\net al. , Applied Physics Letters 109, 032403 (2016)." }, { "title": "1111.3430v1.Frustration_Induced_Ferrimagnetism_in_Heisenberg_Spin_Chains.pdf", "content": "arXiv:1111.3430v1 [cond-mat.str-el] 15 Nov 2011Typeset with jpsj3.cls Full Paper\nFrustration-Induced Ferrimagnetism in Heisenberg Spin Ch ains\nTokuro Shimokawa∗and Hiroki Nakano†\nGraduate School of Material Science, University of Hyogo, K amigori, Hyogo 678-1297, Japan\n(Received October 15, 2019)\nWe study ground-stateproperties of the Heisenberg frustrate d spin chain with interactions\nup to fourth nearest neighbors by the exact-diagonalization meth od and the density matrix\nrenormalizationgroupmethod. We find that ferrimagnetismis realize dnot onlyin the caseof\nS=1/2 but also S=1 despite that there is only a single spin site in each unit cell determine d\nfrom the shape of the Hamiltonian. Our numerical results suggest t hat a “multi-sublattice\nstructure” is not required for the occurrence offerrimagnetism in quantum spin systems with\nisotropic interactions.\nKEYWORDS: quantum spin chain, frustration, ferrimagnetis m, DMRG, exact diagonalization\nFerrimagnetism is a fundamental phenomenon in the field of ma gnetism. One of the most\ntypical examples of ferrimagnetism is the ( S,s) = (1,1/2) mixed spin chain with a nearest-\nneighbor antiferromagnetic (AF) interaction.1)In this system, the so-called Lieb-Mattis-type\nferrimagnetism2,3)is realized in the ground state because two different spins are arranged\nalternately in a line owing to the AF interaction. This syste m includes two spins in a unit\ncell of the system. In other known ferrimagnetic cases of qua ntum spin systems except the\nS= 1/2 Heisenberg frustrated spin chain studied in ref. 4, the sit uation that the system has\nmorespinsthanonein each unit cell hasbeenthesame. Until o urrecent study4)demonstrated\nthe occurrence of ferrimagnetism in the ground state of the S= 1/2 Heisenberg frustrated\nspin chain despite the fact that a unit cell of the chain inclu des only a single spin, namely,\nit has no sublattice structure, it had been unclear whether t he “multi-sublattice structure”\nis required for the occurrence of the ferrimagnetism in a qua ntum spin system composed of\nisotropic interactions. The Hamiltonian examined in ref. 4 is given by\nH=J/summationdisplay\ni[Si·Si+1+1\n2Si·Si+2]\n−J′/summationdisplay\ni[Si·Si+3+1\n2(Si·Si+2+ASi·Si+4)], (1)\nwhere the real constant Ais fixed to be unity. Here, Siis theS= 1/2 spin operator at the\n∗E-mail address: rk09s002@stkt.u-hyogo.ac.jp\n†E-mail address: hnakano@sci.u-hyogo.ac.jp\n1/6J. Phys. Soc. Jpn. Full Paper\nsitei. The numerical study of this system clarified the existence o f the ferrimagnetic ground\nstate when the controllable parameter J′/Jis changed. In addition, research confirmed that\nthere are two types of ferrimagnetic phases: the phase of the Lieb-Mattis (LM) type and the\nphase of the non-Lieb-Mattis (NLM) type, which has been foun d in several frustrated spin\nsystems.5–8)\nThe purpose of this study is to confirm that the above example i s not a special or rare\ncase by investigating other models. In this study, we discus s the ground state of Hamiltonian\n(1) not only in the case of S= 1/2, but also in the case of Sibeing an S= 1 spin operator.\nMoreover, we focus on the case of A= 0.4, which is different from A= 1. Note that energies\nare measured in units of J; we setJ= 1 hereafter.\nWe employ two reliable numerical methods, i.e., the density matrix renormalization group\n(DMRG) method9,10)and the exact-diagonalization (ED) method. Both methods ca n give\nprecise physical quantities for finite-size clusters. The D MRG method is very powerful for\na one-dimensional system under the open-boundary conditio n. On the other hand, the ED\nmethod does not suffer from the limitation posed by the shape of the clusters; there is no\nlimitation of boundary conditions, although the ED method c an treat only systems smaller\nthan those that the DMRG method can treat. Note that, in the pr esent research, we use the\n“finite-system” DMRG method.\nIn the present study, two quantities are calculated. One is t he lowest energy in each\nsubspace divided by Sz\ntotto determine the spontaneous magnetization M, whereSz\ntotis the\nzcomponent of the total spin. We obtain the lowest energy E(N,Sz\ntot,J′) for a system size\nNand a given J′. For example, the Sz\ntotdependence of E(N,Sz\ntot,J′) in a specific case of J′\nis presented in the inset of Fig. 1(a). This inset shows the re sults obtained by our DMRG\ncalculations of the system of N= 72 with the maximum number of retained states ( MS) of\n600, and a number of sweeps ( SW) of 10. One can find the spontaneous magnetization Mfor\nagivenJ′as the highest Sz\ntotamong those at the lowest common energy. (See thearrowhead i n\nthe inset.) The other quantity is the local magnetization in the ground state for investigating\nthe spin structure of the highest- Sz\ntotstate. The local magnetization is obtained by calculating\n/angbracketleftSz\ni/angbracketright, whereSz\niis thez-component of the spin at the site iand/angbracketleftO/angbracketrightdenotes the expectation\nvalue of the physical quantity Owith respect to the state of interest.\nFirst, let us show the results of the J′dependence of M/Msin Fig. 1, where Msis the\nsaturated magnetization. Irrespective of S= 1/2 orS= 1, we find the nonmagnetic phase\n(M/Ms= 0) and ferromagnetic phase( M/Ms= 1). Between thetwo phases, wealso findthree\nregions: the regions of 0 < M/M s<1/3,M/Ms= 1/3, and 1/3< M/M s<1. ForS= 1/2,\n2/6J. Phys. Soc. Jpn. Full Paper\n\tB\n\tC\n 0 5 10 15 Stot z–43–42.5–42E nergy S=1/2 N=72 \nDMR G \nMS=600, SW=10 \nJ’ =2.2 \n0 1 2 \nJ’ 00.51M/M sN=24 DMR G \nN=48 DMR G \nN=72 DMR G S=1 A=0.40 1 2 \nJ’ 00.51M/M s\nN=24 E D periodic\nN=24 E D open\nN=72 DMR G S=1/2 A=0.4\nFig. 1. (Color) (a) J′dependence of the normalized magnetization M/Msin the ground state in the\ncase ofS= 1/2 withA= 0.4. In the inset of (a), the lowest energy in each subspace divided by\nSz\ntotis shown. Results of the DMRG calculations are presented when the s ystem size is N= 72\nforJ′= 2.2. The arrowhead indicates the spontaneous magnetization Mfor a given J′;Mis\ndetermined to be the highest Sz\ntotamong the values with the lowest common energy. (b) J′\ndependence of M/Msin the ground state in the case of S= 1 with A= 0.4.\none can see that the region of 0 < M/M s<1/3 is much narrower than the distinctly existing\nregion of NLM ferrimagnetism4)in the case of S= 1/2 withA= 1. The width of the present\nregion for A= 0.4 seems to vanish in the limit of N→ ∞. One finds that the occurrence\nof the NLM ferrimagnetism in Hamiltonian (1) requires a four th-neighbor interaction with A\nthat is larger than the specific value between A= 0.4 andA= 1. The width of the region of\nM/Ms= 1/3 in both cases of S= 1/2 withA= 0.4 andS= 1 with A= 0.4 seems to survive\nin the limit of N→ ∞. The region of 1 /3< M/M s<1 is presumably considered to merge\nwith the ferromagnetic (FM) phase in the thermodynamic limi t. The reason for this is that\nthis region appears only near M/Ms= 1 and that M/Msin this region becomes progressively\nlarger with increasing N. In addition, we cannot confirm this region in the calculatio ns within\nN≤30 of the S= 1/2 system under the periodic-boundary condition irrespecti ve of the\nvalues of A. The issue of whether or not the region of 1 /3< M/M s<1 survives should be\nclarified in future studies; hereafter, we do not pay further attention to this issue.\n3/6J. Phys. Soc. Jpn. Full Paper\n\tB\n \tC\n0 0.02 0.04\n1/N12J’ J’ 4\nJ’ 3\nJ’ 2\nJ’ 1\n0 0.02 0.04\n1/N00.511.5Width of phase 0J’ =2.1, M=24 \nFig. 3. Local magnetization /angbracketleftSz\ni/angbracketrightunder the open-boundary condition: for J′= 2.1 in the case of\nS= 1 with A= 0.4 from the DMRG calculation for N= 72. The site number is denoted by i,\nwhich is classified into i= 3n−2, 3n−1, and 3n, wherenis an integer. Squares, circles, and\ntriangles mean i= 3n−2, 3n−1, and 3n, respectively.\nNext, we study the size dependences of the phase boundaries i n the case of S= 1 with\nA= 0.4 depicted in Fig. 2(a). We present results of four boundarie s:J′=J′\n1between the\nnonmagnetic phase and the region of 0 < M/M s<1/3,J′=J′\n2between the regions of\n0< M/M s<1/3 andM/Ms= 1/3,J′=J′\n3between the regions of M/Ms= 1/3 and\n1/3< M/M s<1, andJ′=J′\n4between the region of 1 /3< M/M s<1 and the FM phase. To\nconfirm the behavior up to the thermodynamic limit, we also ex amine the N−1dependences\nof the two widths of the regions of M/Ms= 1/3 and 0< M/M s<1/3 in Fig. 2(b). Although\nthe width of the region of M/Ms= 1/3 decreases with increasing N, this dependence shows\na behavior that is convex-downwards for large sizes; the wid th seems to converge to 0.3.\nTherefore, the phase of M/Ms= 1/3 definitely survives in the limit of N→ ∞. On the\nother hand, the width of 0 < M/M s<1/3 obviously disappears in the limit of N→ ∞. An\nappropriate tuning of the parameters in Hamiltonian (1) of t heS= 1 system might cause the\n4/6J. Phys. Soc. Jpn. Full Paper\nNLM ferrimagnetism; such parameter sets should be searched for in future studies.\nFinally, we examine the local magnetization /angbracketleftSz\ni/angbracketrightin the phase of M/Ms= 1/3 in the\ncase ofS= 1 with A= 0.4. In Fig. 3, we present our DMRG result of /angbracketleftSz\ni/angbracketrightof the system\nofN= 72. We confirm the up-down-up spin behavior, and this spin st ructure is consistent\nwithM/Ms=1/3 in the parameter region near approximately J′= 2.1 in Fig. 1(b). Thus, this\nphase is considered to be the LM-type ferrimagnetic phase.\nIn summary, we study the ground-state properties of a frustr ated Heisenberg spin chain by\nthe ED and DMRG methods. Despite the fact that this system con sists of only a single spin\nsiteineach unitcell determinedfromtheshapeof theHamilt onian, theLM-typeferrimagnetic\nground state is realized in a finite region not only in the case ofS= 1/2 but also of S= 1.\nThe present models showing ferrimagnetism indicate that a “ multi-sublattice structure” is\nnot required for the occurrence of ferrimagnetism in quantu m spin systems with isotropic\ninteractions as a general circumstance.\nWe are grateful to Professor Y. Hasegawa for his critical rea ding of the manuscript. This\nwork was partly supported by Grants-in-Aid (Nos. 20340096, 23340109, and 23540388) from\nthe Ministry of Education, Culture, Sports, Science and Tec hnology (MEXT) of Japan. This\nwork was partly supported by a Grant-in-Aid (No. 22014012) f or Scientific Research and\nPriority Areas “Novel States of Matter Induced by Frustrati on” from the MEXT of Japan.\nSome of the calculations were carried out at the Supercomput er Center, Institute for Solid\nState Physics, University of Tokyo. Exact-diagonalizatio n calculations in the present work\nwere carried out based on TITPACK Version 2 coded by H. Nishim ori. DMRG calculations\nwere carried out using the ALPS DMRG application.11)\n5/6J. Phys. Soc. Jpn. Full Paper\nReferences\n1) T. Sakai and K. Okamoto: Phys. Rev. B. 65(2002) 214403.\n2) E. Lieb and D. Mattis: J. Math. Phys. 3(1962) 749.\n3) W. Marshall: Proc. Roy. Soc. A 232(1955) 48.\n4) T. Shimokawa and H. Nakano: J. Phys. Soc. Jpn. 80(2011) 043703.\n5) S. Yoshikawa and S. Miyashita: J. Phys. Soc. Jpn. 74(2005) Suppl. 71.\n6) K. Hida: J. Phys.: Condens. Matter 19(2007) 145225.\n7) H. Nakano, T. Shimokawa, and T. Sakai: J. Phys. Soc. Jpn. 80(2011) 033709.\n8) T. Shimokawa and H. Nakano: J. Phys.: Conf. Ser. 320(2011) 012007.\n9) S. R. White: Phys. Rev. Lett. 69(1992) 2863.\n10) S. R. White: Phys. Rev. B. 48(1993) 10345.\n11) A. F. Albuquerque, F. Alet, P. Corboz, P. Dayal, A. Feiguin, L. G amper, E. Gull, S. Gurtler, A.\nHonecker, R. Igarashi, M. Korner, A. Kozhevnikov, A. Lauchli, S. R. Manmana, M. Matsumoto,\nI. P. McCulloch, F. Michel, R. M. Noack, G. Pawlowski, L. Pollet, T. Pru schke, U. Schollwock, S.\nTodo, S. Trebst, M. Troyer, P. Werner, and S. Wessel: J. Magn. M agn. Mater. 310(2007) 1187\n(see also http://alps.comp-phys.org).\n6/6" }, { "title": "2204.14010v4.Entangling_mechanical_vibrations_of_two_massive_ferrimagnets_by_fully_exploiting_the_nonlinearity_of_magnetostriction.pdf", "content": "Entangling mechanical vibrations of two massive ferrimagnets by fully exploiting\nthe nonlinearity of magnetostriction\nHang Qian,1Zhi-Yuan Fan,1and Jie Li1,\u0003\n1Interdisciplinary Center of Quantum Information, State Key Laboratory of Modern Optical Instrumentation,\nand Zhejiang Province Key Laboratory of Quantum Technology and Device,\nDepartment of Physics, Zhejiang University, Hangzhou 310027, China\n(Dated: December 14, 2022)\nQuantum entanglement in the motion of macroscopic objects is of significance to both fundamental studies\nand quantum technologies. Here we show how to entangle the mechanical vibration modes of two massive\nferrimagnets that are placed in the same microwave cavity. Each ferrimagnet supports a magnon mode and a\nlow-frequency vibration mode coupled by the magnetostrictive force. The two magnon modes are, respectively,\ncoupled to the microwave cavity by the magnetic dipole interaction. We first generate a stationary nonlocal\nentangled state between the vibration mode of the ferrimagnet-1 and the magnon mode of the ferrimagnet-2.\nThis is realized by continuously driving the ferrimagnet-1 with a strong red-detuned microwave field and the en-\ntanglement is achieved by exploiting the magnomechanical parametric down-conversion and the cavity-magnon\nstate-swap interaction. We then switch o \u000bthe pump on the ferrimagnet-1 and, simultaneously, turn on a red-\ndetuned pulsed drive on the ferrimagnet-2. The latter drive is used to activate the magnomechanical beamsplitter\ninteraction, which swaps the magnonic and mechanical states of the ferrimagnet-2. Consequently, the previously\ngenerated phonon-magnon entanglement is transferred to the mechanical modes of two ferrimagnets. The work\nprovides a scheme to prepare entangled states of mechanical motion of two massive objects, which may find\napplications in various studies exploiting macroscopic entangled states.\nI. INTRODUCTION\nEntanglement of mechanical motion has been first demon-\nstrated in two microscopic trapped atomic ions [1], then in two\nsingle-phonon excitations in nanodiamonds [2], and recently\nin two macroscopic optomechanical resonators [3–5]. In the\npast two decades, a number of theoretical proposals have been\ngiven in optomechanics [6–35] for preparing the entanglement\nbetween massive mechanical resonators, utilizing the cou-\npling between optical and mechanical degrees of freedom by\nradiation pressure. In particular, the application of reservoir\nengineering ideas [36–40] to optomechanics [15, 16, 20, 22]\nhas led to a significant and robust mechanical entanglement\nand the entanglement has been successfully demonstrated in\nthe experiment [4].\nExploring novel physical platforms that could prepare\nquantum states at a more massive scale is of great signifi-\ncance for the study of macroscopic quantum phenomena [41],\nthe boundary between the quantum and classical worlds [42],\nand gravitational quantum physics [43, 44], etc. Recently, the\nmagnomechanical system of a large-size ferrimagnet, e.g., yt-\ntrium iron garnet (YIG), that has a dispersive magnon-phonon\ncoupling has shown such a potential [45–58]. The magnon\nand mechanical vibration modes are coupled by the nonlin-\near magnetostrictive interaction, which couples ferrimagnetic\nmagnons to the deformation displacement of the ferrimag-\nnet [59–62]. The magnomechanical Hamiltonian takes the\nsame form as the optomechanical one [63] (by exchanging\nthe roles of magnons and photons), which allows us to pre-\ndict many optomechanical analogues in magnomechanics. By\ncoupling the magnomechanical system to a microwave cav-\n\u0003jieli007@zju.edu.cnity, they form the tripartite cavity magnomechanical system.\nThe magnetostriction has been exploited in cavity magnome-\nchanics to generate macroscopic entangled states of magnons\nand vibration phonons of massive ferrimagnets [45, 47, 48,\n51, 52, 56, 58], as well as nonclassical states of microwave\nfields [64, 65]. Therefore, the magnetostrictive nonlinearity\nbecomes a valuable resource for producing various quantum\nstates of microwave photons, magnons, and phonons. These\nnonclassical states may find potential applications in quan-\ntum information processing [66, 67], quantum metrology [65]\nand quantum networks [68]. Despite of the aforementioned\nmany proposals and very limited experiments [60–62] in cav-\nity magnomechanics, there is so far only one protocol [51] for\nentangling two mechanical vibration modes of macroscopic\nferrimagnets. The protocol [51], however, relies on an exter-\nnal entangled resource and transfers the quantum correlation\nfrom microwave drive fields to two vibration modes. There-\nfore, designing more energy-saving protocols without using\nany external quantum resource is highly needed.\nAlong this line, we present here a scheme for generating\na nonlocal entangled state between the mechanical vibrations\nof two ferrimagnets, without the need of any quantum driving\nfield. Two ferrimagnets are placed in a microwave cavity and\neach ferrimagnet supports a magnon mode and a mechanical\nmode. The cavity mode couples to two magnon modes by\nthe magnetic dipole interaction, and the magnon modes cou-\nple to their local vibration modes by magnetostriction, respec-\ntively. We show that the entanglement between the vibration\nmodes of two ferrimagnets can be achieved by fully exploit-\ning the nonlinear magnetostriction interaction (i.e., exploit-\ning both the magnomechanical parametric down-conversion\n(PDC) and state-swap interactions) and by using the common\ncavity field being an intermediary to distribute quantum cor-\nrelations. The mechanical entanglement is established by two\nsteps. We first generate steady-state entanglement between thearXiv:2204.14010v4 [quant-ph] 13 Dec 20222\nmechanical mode of the ferrimagnet-1 and the magnon mode\nof the ferrimagnet-2. We then activate the magnomechani-\ncal state-swap interaction in the ferrimagnet-2, which trans-\nfers the magnonic state to its locally interacting mechanical\nmode. Consequently, the two mechanical modes get nonlo-\ncally entangled.\nThe remainder of the paper is organized as follows. In\nSec. II, we introduce the general model of the protocol that is\nused in the two steps. We then show how to prepare a station-\nary nonlocal magnon-phonon entanglement with a continuous\nmicrowave pump in Sec. III, and how to transfer this entan-\nglement to two mechanical modes with a pulsed microwave\ndrive in Sec. IV. Finally, we discuss and conclude in Sec. V.\nII. THE MODEL\nThe protocol is based on a hybrid five-mode cavity mag-\nnomechanical system, including a microwave cavity mode,\ntwo magnon modes, and two mechanical vibration modes,\nas depicted in Fig. 1. The magnon modes are embodied by\nthe collective motion of a large number of spins (i.e., spin\nwave) in two ferrimagnets, e.g., two YIG spheres [60–62] or\nmicro bridges [69, 70]. They simultaneously couple to the mi-\ncrowave cavity via the magnetic dipole interaction. This cou-\npling can be strong thanks to the high spin density in YIG [71–\n73]. The mechanical modes refer to the deformation vibration\nmodes of two YIG crystals caused by the magnetostrictive\nforce. Due to the much lower mechanical frequency (rang-\ning from 100to 102MHz) than the magnon frequency (GHz)\nin the typical magnomechanical systems [60–62, 69], the vi-\nbration phonons and the magnons are coupled in a dispersive\nmanner [70, 74, 75]. The Hamiltonian of the system reads\nH=~=!aaya+X\nj=1;2\u0012\n!mjmy\njmj+!bj\n2\u0010\np2\nj+q2\nj\u0011\u0013\n+X\nj=1;2\u0012\ngj\u0010\namy\nj+aymj\u0011\n+G0jmy\njmjqj\u0013\n+i\nk\u0010\nmy\nke\u0000i!0kt\u0000mkei!0kt\u0011\n:(1)\nThe first (second) term describes the energy of the cavity\nmode (magnon modes), of which the frequency is !a(!mj)\nand the annihilation operator is a(mj) with the commutation\nrelation [ a;ay]=1\u0010\u0002mj;my\nj\u0003=1\u0011\n. The magnon frequency\n!mjis determined by the external bias magnetic field Hjvia\n!mj=\rHj, where the gyromagnetic ratio \r=2\u0019=28 GHz=T.\nThe third term denotes the energy of two mechanical vibra-\ntion modes with frequencies !bj, and qjandpj([qj;pj]=i)\nare the dimensionless position and momentum of the vibra-\ntion mode j, modeled as a mechanical oscillator. The cou-\npling gjis the linear cavity-magnon coupling rate, and G0j\nis the bare magnomechanical coupling rate. For large-size\nYIG spheres with the diameter in the 100 \u0016m range [60–\n62], G0jis typically in the 10 mHz range [75]. It, however,\ncan be much stronger for micron-sized YIG bridges [69, 70].\nNevertheless, the e \u000bective magnomechanical coupling can be\nsignificantly enhanced by driving the magnon mode with a\nFIG. 1: (a) Sketch of the system. Two YIG crystals are placed near\nthe maximum magnetic fields of a microwave cavity. Each YIG crys-\ntal is in a uniform bias magnetic field, and supports a magnon mode\nand a mechanical vibration mode. The two magnon modes couple to\nthe same cavity field. Two drive fields are applied successively to the\ntwo magnon modes in the two steps of the protocol. Note that though\nYIG spheres are adopted in the sketch, the YIG crystals can be a\nnonspherical structure, e.g., micron-sized YIG bridges [69, 70]. (b)\nInteractions among the subsystems. The magnon mode mj(j=1;2)\ncouples linearly to the cavity mode awith the coupling strength gj,\nand couples dispersively to the mechanical mode bjwith the e \u000bective\nmagnomechanical coupling rate Gj. A nonlocal phonon-magnon ( b1-\nm2) entanglement is created in the first step with a continuous drive\non the magnon mode m1, and the entanglement is transferred to the\ntwo mechanical modes in the second step by using a pulsed drive on\nthe magnon mode m2.\nstrong microwave field [45]. The driving Hamiltonian is de-\nscribed by the last term, and the corresponding Rabi frequency\n\nk=p\n5\n4\rpNkBk(k=1 or 2) [45], with Bk(!0k) being\nthe amplitude (frequency) of the drive magnetic field, and Nk\nbeing the total number of spins in the kth crystal. We re-\nmark that the model di \u000bers from the one used in Ref. [47]\nby including a second mechanical mode, which brings in a\nsignificant amount of additional thermal noise to the system.\nMore importantly, the present work aims to entangle two low-\nfrequency (in MHz) mechanical modes. This is much more\ndi\u000ecult to prepare than the entanglement of two GHz magnon\nmodes studied in Ref. [47].\nIn what follows, we adopt a two-step procedure to prepare\nthe two mechanical modes in an entangled state, and in each\nstep, we apply a single drive field on either magnon mode\nm1orm2. This avoids the complex Floquet dynamics in our\nhighly hybrid system caused by simultaneously applying mul-\ntiple pump tones [15, 16, 20, 22]. We first generate a nonlo-\ncal entangled state between the mechanical mode b1and the3\nFIG. 2: Mode and drive frequencies used in the first step. When\nthe magnon mode m1is (the cavity and magnon mode m2are) reso-\nnant with the blue (red) mechanical sideband of the drive field with\nfrequency!01, the nonlocal phonon-magnon entanglement Eb1m2is\nestablished.\nmagnon mode m2by continuously driving the magnon mode\nm1. After the system enters a stationary state, we then turn\no\u000bthe drive on m1and, simultaneously, turn on a red-detuned\ndrive on the magnon mode m2to activate the magnomechan-\nical state-swap interaction m2$b2. This operation transfers\nthe quantum correlation shared between b1andm2to two me-\nchanical modes, thus establishing a quantum correlation (i.e.,\nentanglement) between the two mechanical modes.\nIII. STATIONARY NONLOCAL MAGNO-MECHANICAL\nENTANGLEMENT\nIn the first step, we aim to entangle the mechanical mode\nb1and the magnon mode m2. This can be realized by driving\nthe magnon mode m1with a strong red-detuned microwave\nfield [45], see Fig. 2. It was shown that a genuine tripar-\ntite magnon-photon-phonon entangled state can be produced\nwithout involving the second YIG crystal [45]. By using the\npartial result that the mechanical mode b1and the cavity a\nare entangled, and by coupling the cavity to the second nearly\nresonant magnon mode m2(which have a beamsplitter inter-\naction realizing the state-swap operation a$m2), the two\nmodes b1andm2thus get entangled. This is confirmed by the\nnumerical results presented in this section.\nIt should be noted that since the strong drive is applied on\nthefirst YIG crystal, the e \u000bective (magnomechanical) cou-\npling to the mechanical mode b2in the second YIG crystal is\nmuch smaller than that in the first YIG crystal, G2\u001cG1, so\nthe presence of the second mechanical mode, or not, will not\nappreciably a \u000bect the entanglement dynamics analysed above.\nBecause in the next step, the coupling to the second mechani-\ncal mode must be turned on ( G02>0), including this coupling\nin the model also in the first step means no additional opera-\ntion (e.g., adjusting the direction of the bias magnetic field\nto activate or inactivate the coupling G02[47, 60]) has to be\nimplemented between the two steps. Another reason is that,\nas will be shown in Sec. IV, the entanglement between b1-m2\nshould be transferred to the two mechanical modes as soon as\npossible because it rapidly decays when the drive in the first\nstep is switched o \u000b.\nIn the frame rotating at the drive frequency !01, the quan-\ntum Langevin equations (QLEs) describing the system dy-namics are given by\n˙a=\u0000(i\u0001a+\u0014a)a\u0000ig1m1\u0000ig2m2+p\n2\u0014aain;\n˙mj=\u0000(i\u0001mj+\u0014mj)mj\u0000iG0jmjqj\u0000igja+ \n j+q\n2\u0014mjmin\nj;\n˙qj=!bjpj;\n˙pj=\u0000!bjqj\u0000\rbjpj\u0000G0jmy\njmj+\u0018j;\n(2)\nwhere \u0001a=!a\u0000!01,\u0001mj=!mj\u0000!01, and\u0014a,\u0014mjand\rbj\nare the dissipation rates of the cavity, magnon and mechanical\nmodes, respectively. The Rabi frequency \nj= \n 1\u000ej1(j=\n1;2) implies only one drive field applied on the magnon mode\nm1.ainandmin\njare the input noise operators a \u000becting the cav-\nity and magnon modes, whose non-zero correlation functions\narehain(t)ainy(t0)i=[Na(!a)+1]\u000e(t\u0000t0),hainy(t)ain(t0)i=\nNa(!a)\u000e(t\u0000t0),hmin\nj(t)miny\nj(t0)i=[Nmj(!mj)+1]\u000e(t\u0000t0)\nandhminy\nj(t)min\nj(t0)i=Nmj(!mj)\u000e(t\u0000t0). The Langevin force\noperator\u0018jis accounting for the mechanical Brownian mo-\ntion, which is autocorrelated as h\u0018j(t)\u0018j(t0)+\u0018j(t0)\u0018j(t)i \u0019\n\rbj[2Nbj(!bj)+1]\u000e(t\u0000t0), where we consider a high qual-\nity factor Qb=!b=\rb\u001d1 for the mechanical oscillators\nto validate the Markov approximation [76]. Here, Nk(!k)=h\nexp(~!k\nkBT)\u00001i\u00001(k=a;mj;bj) are the equilibrium mean ther-\nmal photon, magnon, and phonon numbers, respectively, at\nthe environmental temperature T, with kBbeing the Boltz-\nmann constant.\nThe strong drive field leads to large amplitudes of the\nmagnon modes and cavity mode due to the magnon-cavity\ncoupling,jhmjij;jhaij \u001d 1. This allows us to linearize the\nnonlinear QLEs (2) by writing each mode operator as a classi-\ncal average plus a fluctuation operator with zero mean value,\ni.e., O=hOi+\u000eO(O=a;mj;qj;pj), and by neglect-\ning small second-order fluctuation terms. Substituting the\nabove mode operators into Eq. (2) yields two sets of lin-\nearized Langevin equations, respectively, for classical aver-\nages and fluctuation operators. By solving the former set of\nequations in the time scale where the system evolves into a\nstationary state, we obtain the solution of the steady-state av-\nerages, which are hpji=0,hqji=\u0000G0jjhmjij2=!bj,hai=\n\u0000i\u0000g1hm1i+g2hm2i\u0001=(i\u0001a+\u0014a), and\nhm1i=\n1(i\u0001a+\u0014a)\ng2\n1+(i˜\u0001m1+\u0014m1)(i\u0001a+\u0014a)\u0000g2\n1g2\n2\ng2\n2+(i˜\u0001m2+\u0014m2)(i\u0001a+\u0014a);\nhm2i=\u0000\n1(i\u0001a+\u0014a)g1g2\ng2\n1+(i˜\u0001m1+\u0014m1)(i\u0001a+\u0014a)\ng2\n2+(i˜\u0001m2+\u0014m2)(i\u0001a+\u0014a)\u0000g2\n1g2\n2\ng2\n1+(i˜\u0001m1+\u0014m1)(i\u0001a+\u0014a);\n(3)\nwhere ˜\u0001mj= \u0001 mj+G0jhqjiis the e \u000bective magnon-drive\ndetuning, which includes the magnetostriction induced fre-\nquency shift. Typically, this frequency shift is negligible (be-\ncause of a small G0[60–62]) with respect to the optimal de-\ntuning used in this work, i.e., j˜\u0001mj\u0000\u0001mjj \u001c j \u0001mjj \u0019!bj.\nTherefore, in what follows we can safely assume ˜\u0001mj\u0019\u0001mj.4\nThe set of the linearized QLEs for the system fluctuations\ncan be written in the matrix form\n˙u(t)=Au(t)+n(t); (4)\nwhere u(t)=\u0002\u000eX(t);\u000eY(t);\u000ex1(t);\u000ey1(t);\u000ex2(t);\u000ey2(t);\u000eq1(t);\n\u000ep1(t);\u000eq2(t);\u000ep2(t)\u0003Tis the vector of the quadrature fluc-\ntuations, and n(t)=hp2\u0014aXin(t);p2\u0014aYin(t);p\n2\u0014m1xin\n1(t);p\n2\u0014m1yin\n1(t);p\n2\u0014m2xin\n2(t);p\n2\u0014m2yin\n2(t);0; \u0018 1(t);0; \u0018 2(t)iT\nis the vector of the input noises entering the system,\nwith\u000eX=\u0010\n\u000ea+\u000eay\u0011\n=p\n2,\u000eY=i\u0010\n\u000eay\u0000\u000ea\u0011\n=p\n2,\n\u000exj=\u0010\n\u000emj+\u000emy\nj\u0011\n=p\n2, and\u000eyj=i\u0010\n\u000emy\nj\u0000\u000emj\u0011\n=p\n2,\nand the drift matrix Ais given by\nA=0BBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBBB@\u0000\u0014a\u0001a 0 g1 0 g2 0 0 0 0\n\u0000\u0001a\u0000\u0014a\u0000g1 0\u0000g2 0 0 0 0 0\n0 g1\u0000km1 \u0001m1 0 0\u0000Re[G1]0 0 0\n\u0000g10\u0000\u0001m1\u0000\u0014m1 0 0\u0000Im[G1]0 0 0\n0 g2 0 0 \u0000\u0014m2 \u0001m2 0 0\u0000Re[G2]0\n\u0000g20 0 0 \u0000\u0001m2\u0000\u0014m2 0 0\u0000Im[G2]0\n0 0 0 0 0 0 0 !b1 0 0\n0 0\u0000Im[G1]Re[G1] 0 0\u0000!b1\u0000\rb1 0 0\n0 0 0 0 0 0 0 0 0 !b2\n0 0 0 0 \u0000Im[G2]Re[G2] 0 0\u0000!b2\u0000\rb21CCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCA; (5)\nwith Gj=ip\n2G0jhmjibeing the e \u000bective magnomechanical\ncoupling rate, which is generally complex.\nBecause of the linearized dynamics of the system and the\nGaussian nature of the input noises, the steady state of the\nsystem’s quantum fluctuations is a five-mode Gaussian state.\nThe state can be completely described by a 10 \u000210 covariance\nmatrix (CM)Vwith entries defined as Vi j=hui(t)uj(t0)+\nuj(t0)ui(t)i=2 (i;j=1;2;:::; 10). The CMVcan be obtained\nby directly solving the Lyapunov equation\nAV+VAT=\u0000D; (6)\nwhere the di \u000busion matrix Dis defined as Di j=\nhni(t)nj(t0)+nj(t0)ni(t)i=\u00022\u000e(t\u0000t0)\u0003and given by D =\ndiag[\u0014a(2Na+1);\u0014a(2Na+1);\u0014m1(2Nm1+1);\u0014m1(2Nm1+1);\n\u0014m2(2Nm2+1);\u0014m2(2Nm2+1);0;\rb1(2Nb1+1);0;\rb2(2Nb2+1)].\nWhen the CM of the system fluctuations is obtained, one\ncan then extract the state of the interesting modes by remov-\ning inVthe rows and columns related to the uninteresting\nmodes and study their entanglement properties. We use the\nlogarithmic negativity [77] to quantify the Gaussian bipartite\nentanglement, of which the definition is provided in the Ap-\npendix.\nIn Fig. 3, we show relevant steady-state bipartite entan-\nglements generated in the first step. The blank areas are\nthe parameter regimes in which the system is unstable, re-\nflected by the fact that the drift matrix Ahas at least one pos-\nitive eigenvalue. We have employed the following parame-\nters [60–62, 69, 70]: !a=2\u0019=10 GHz,!b1=2\u0019=17 MHz,\n!b2=2\u0019=12 MHz,\rb=2\u0019=100 Hz,\u0014a=2\u0019=1 MHz,\u0014m=\u0014a,\ng1=2\u0019=5 MHz, g2=2\u0019=1 MHz, G01=2\u0019'G02=2\u0019=10\nHz, and T=10 mK. The amplitude of the drive magnetic\nfield B1=4:8\u000210\u00004T, corresponding to the drive power\nP1'1:1 mW for two YIG micro bridges (approximated as\nFIG. 3: Nonlocal phonon-magnon entanglement Eb1m2versus (a) \u0001a\nand\u0001m1, and (b) \u0001aand\u0001am2\u0011!a\u0000!m2. (c) The phonon-cavity\nentanglement Eb1aand (d) the mechanical entanglement Eb1b2versus\n\u0001aand\u0001m1. In (d) the whole purple region denotes Eb1b2=0. We\ntake\u0001am2=0:9\u0014ain (a), (c) and (d), and \u0001m1=0:95!b1in (b). The\nblank areas denote the unstable region where the stability condition\nis not fulfilled. See the text for other parameters.\ntwo cuboids) with dimensions of 13.7 \u00023\u00021\u0016m3and 16.4\n\u00023\u00021\u0016m3. This milliwatt power might lead to heating in-\nduced temperature rise. To avoid the possible heating e \u000bect,\none should improve the bare coupling G0jto reduce the power.\nSince the mechanical resonance frequency mainly depends on\nthe length and thickness of the beam [69], the sizes of the\nbeams are chosen corresponding to the mechanical frequen-5\ncies we adopted. To determine the power, we have used the\nrelation between the drive magnetic field B1and the power P1,\ni.e.,B1=p\n2\u00160P1=(lwc) [70], with \u00160being the vacuum mag-\nnetic permeability, cthe speed of the electromagnetic wave\npropagating in vacuum, and land wthe length and width\nof the YIG cuboid. The nonlocal phonon-magnon entangle-\nment Eb1m2(Fig. 3(a)-(b)) is the result of the phonon-cavity\nentanglement Eb1a(Fig. 3(c)) and the cavity-magnon ( a-m2)\nbeamsplitter (state-swap) interaction. The phonon-cavity en-\ntanglement Eb1ais obtained by using the results of Ref. [45],\nwhich studies a tripartite magnon-photon-phonon system (i.e.,\nthem1-a-b1subsystem here). As shown in Ref. [45], a red-\ndetuned strong drive field on the magnon mode m1(with\n\u0001m1\u0019!b1) cools the hot mechanical mode b1(a precondi-\ntion for preparing quantum states) and, simultaneously, acti-\nvates the magnomechanical PDC interaction, which produces\nthe magnomechanical entanglement Eb1m1. The latter process\noccurs only when the pump field is su \u000eciently strong, such\nthat the weak coupling condition jG1j\u001c!b1for taking the\nrotating-wave (RW) approximation to obtain the cooling in-\nteraction/my\n1b1+m1by\n1is no longer satisfied, and the counter-\nRW terms/my\n1by\n1+m1b1(responsible for the PDC interac-\ntion) start to play the role. The magnomechanical entangle-\nment is partially transferred to the phonon-cavity subsystem\nwith Eb1a>0 (Fig. 3(c)) when \u0001a\u0019\u0000!b1[45]. The opti-\nmal detunings \u0001m1\u0019 \u0000\u0001a\u0019!b1imply that the cavity and\nthe magnon mode m1are respectively resonant with the two\nmechanical sidebands of the drive field, see Fig. 2.\nSince the cavity-magnon ( m2) state-swap operation works\noptimally when the two modes are resonant. The entangle-\nment Eb1m2is maximized for a nearly zero cavity-magnon de-\ntuning \u0001am2\u00190, as confirmed by Fig. 3(b). The detuning\n\u0001am2up to several cavity linewidth will significantly hinder\nthe transfer of the entanglement. The fact that the entangle-\nment Eb1m2is transferred from the entanglement Eb1acan also\nbe seen from the complementary distribution of the entangle-\nment in Figs. 3(a) and 3(c), indicating the entanglement flow\namong the subsystems switched on by the beamsplitter cou-\npling.\nIt is worth noting that in the first step the two mechani-\ncal modes are not entangled, as confirmed by Eb1b2=0 in\nFig. 3(d) in a wide range of parameters. This is mainly be-\ncause the magnon mode m2is driven by the nearly resonant\ncavity field, which is not the condition for cooling the me-\nchanical mode b2, or realizing the state-swap interaction be-\ntween the two modes m2andb2. Instead, a red-detuned mi-\ncrowave field should be used to drive the magnon mode m2\nto realize the state-swap operation m2$b2, such that the\nphonon-magnon entanglement Eb1m2can be transferred to the\ntwo mechanical modes, as will be discussed in the next sec-\ntion.\nIV . ENTANGLEMENT BETWEEN TWO MECHANICAL\nMODES\nIn this section, we show how to transfer the phonon-\nmagnon entanglement Eb1m2prepared in the first step to the\nFIG. 4: Mode and drive frequencies used in the second step. A red-\ndetuned pulsed drive with frequency !02is used to activate the mag-\nnomechanical state-swap operation m2$b2, which transfers the\nphonon-magnon entanglement Eb1m2generated in the first step to the\ntwo mechanical modes.\ntwo mechanical modes. To this end, we apply a red-detuned\npulsed microwave drive on the magnon mode m2(see Fig. 4)\nto activate the local magnomechanical state-swap interaction\nm2$b2. The pulsed drive should be fast as the entangle-\nment Eb1m2will quickly decay as soon as the continuous drive\nis turned o \u000bin the first step. To simplify the model, we use\na flattop microwave pulse to drive the magnon mode and then\nthe model and the treatment used in Sec. III for a continuous\ndrive are still valid for the case of a flattop pulse drive. Di \u000ber-\nently, we shall solve the dynamical solutions rather than the\nsteady-state solutions.\nThe QLEs remain the same as in Eq. (2), except that all\nthe detunings are redefined with respect to the frequency !02\nof the pulsed drive and the Rabi frequency is associated with\nthe pulsed drive, i.e., \u0001a=!a\u0000!02,\u0001mj=!mj\u0000!02, and\n\nj= \n 2\u000ej2. The dynamics of the quantum fluctuations of the\nsystem can still be described by Eq. (4) but with dynamical\nhmji(t) and thus Gj(t) in the drift matrix, due to a pulsed drive.\nThe dynamical CM V(t) can be obtained by [29]\nV(t)=M(t)V0M(t)T+Zt\n0ds M (s)DM(s)T; (7)\nwhere tis the duration of the pulsed drive, M(t)=eRt\n0A(\u001c)d\u001c,\nandV0is the CM of the initial state of the system when si-\nmultaneously turn on (o \u000b) the drive on the magnon mode m2\n(m1), which is obtained in the first step by solving the Lya-\npunov equation (6). When the dynamical CM is achieved, we\ncan then study the dynamics of the entanglement.\nFigure 5(a) shows the mechanical entanglement Eb1b2is\nmaximized at the detuning \u0001m2\u0019!b2, which is the optimal\ndetuning for realizing the magnomechanical state-swap inter-\naction, by which the previously generated phonon-magnon\nentanglement Eb1m2is transferred to the mechanical modes.\nNote that Fig. 5(a) is plotted at the optimal pulse duration tmax\nfor a given detuning, yielding a maximal Eb1b2. Figure 5(b)\nshows Eb1b2versus the pulse duration tfor the optimal detun-\ning. As is shown, there is a time window for the presence of\nthe entanglement. The mechanical entanglement emerges a\nshort while after the phonon-magnon entanglement dies out\n(c.f. the inset of Fig. 5(b)). When Eb1b2reaching its maxi-\nmum at tmax, we then turn o \u000bthe pulse drive to decouple the\nmechanics from the rest of the system to protect the entan-\nglement. The two mechanical oscillators then evolve almost\nfreely, and their entanglement is a \u000bected by their local thermal\nbaths, which lasts for a much longer time due to a small me-6\nFIG. 5: Mechanical entanglement Eb1b2versus (a) detuning \u0001m2at\nthe optimal pulse duration tmax; (b) pulse duration tat the optimal\ndetuning \u0001m2=!b2. The inset shows that the phonon-magnon en-\ntanglement Eb1m2dies out around t'0:01\u0016s. (c) Eb1b2versus time\nwhen the drive is turned o \u000battmax. Soon after tmax, the two mechani-\ncal oscillators evolve freely with the only coupling to their local ther-\nmal baths. (d) Eb1b2versus bath temperature at the optimal detuning\nand pulse duration. We take \u0001a=\u00000:95!b1,\u0001m1=0:95!b1(as\ndefined in the first step), and \u0001am2=0:9\u0014ain all plots. The other\nparameters are the same as in Fig. 3.\nchanical damping and a low bath temperature. This is clearly\nshown in Fig. 5(c) (c.f. Fig. 5(b)). The mechanical entan-\nglement is robust against bath temperature and the maximal\nentanglement survives up to T'100 mK, as illustrated in\nFig. 5(d). We have used a drive power P2=1:3 mW, giving\nthe amplitude of the drive magnetic field B2=4:8\u000210\u00004T.\nLastly, we discuss how to detect the mechanical entangle-\nment. The entanglement can be verified by measuring the CM\nof the two mechanical modes. The mechanical quadratures\ncan be measured by coupling the deformation displacement to\nan optical cavity that is driven by a weak red-detuned light to\ntransfer the mechanical state to the optical field. By homo-\ndyning two cavity output fields one can then obtain the CM of\nthe mechanical modes [5, 78].V . CONCLUSION AND DISCUSSION\nWe present a protocol to entangle two mechanical vibra-\ntion modes in a cavity magnomechanical system. The pro-\ntocol contains two steps by applying successively two drive\nfields on two magnon modes to activate di \u000berent functions of\nthe nonlinear magnetostrictive interaction, namely, the mag-\nnomechanical PDC and state-swap operations. We show that\nthe entanglement between two mechanical vibration modes of\ntwo YIG crystals can be achieved by fully exploiting the above\nmagnetostrictive functions and the cavity-magnon state-swap\ninteraction. We remark that our protocol is valid for any mag-\nnomechanical system of ferrimagnets or ferromagnets, spher-\nical [60–62] or nonspherical structures [69, 70], as long as\nthey possess the dispersive coupling between magnons and\nphonons. The work may find important applications in many\nstudies that require the preparation of macroscopic entangled\nstates.\nAcknowledgments\nThis work has been supported by National Key Re-\nsearch and Development Program of China (Grant No.\n2022YFA1405200) and National Natural Science Foundation\nof China (Nos. 92265202 and 11874249).\nAppendix\nThe logarithmic negativity is used to quantify the Gaussian\nbipartite entanglement, which is defined as\nEN\u0011max\u00020;\u0000ln 2˜\u0017\u0000\u0003; (8)\nwhere ˜\u0017\u0000=min eigji\n2˜V4j(\n2=\b2\nj=1i\u001byand\u001byis the y-\nPauli matrix) is the minimum symplectic eigenvalue of the\nCM ˜V4=PV 4P, withV4being the 4\u00024 CM of two relevant\nmodes, obtained by removing in Vthe rows and columns of\nthe uninteresting modes, and P=diag(1;\u00001;1;1) being the\nmatrix that implements the partial transposition of the CM.\n[1] J. D. Jost, J. P. Home, J. M. Amini, D. Hanneke, R. Ozeri, C.\nLanger, J. J. Bollinger, D. Leibfried, and D. J. Wineland, Nature\n(London) 459, 683 (2009).\n[2] K. C. Lee, M. R. Sprague, B. J. Sussman, J. Nunn, N. K. Lang-\nford, X.-M. Jin, T. Champion, P. Michelberger, K. F. Reim, D.\nEngland, D. Jaksch, and I. A. Walmsley, Science 334, 1253\n(2011).\n[3] R. Riedinger, A. Wallucks, I. Marinkovic, C. L ¨oschnauer, M.\nAspelmeyer, S. Hong, and S. Gr ¨oblacher, Nature (London) 556,\n473 (2018).\n[4] C. F. Ockeloen-Korppi, E. Damsk ¨agg, J.-M. Pirkkalainen, M.\nAsjad, A. A. Clerk, F. Massel, M. J. Woolley, M. A. Sillanp ¨a¨a,\nNature (London) 556, 478 (2018).[5] K. Kotler, G. A. Peterson, E. Shojaee, F. Lecocq, K. Cicak, A.\nKwiatkowski, S. Geller, S. Glancy, E. Knill, R. W. Simmonds,\nJ. Aumentado, and J. D. Teufel, Science 372, 622 (2021).\n[6] S. Mancini, V . Giovannetti, D. Vitali and P. Tombesi, Phys. Rev.\nLett. 88, 120401 (2002).\n[7] J. Zhang, K. C. Peng, and S. L. Braunstein, Phys. Rev. A 68,\n013808 (2003).\n[8] M. Pinard, A. Dantan, D. Vitali, O. Arcizet, T. Briant and A.\nHeidmann, Europhys. Lett. 72, 747 (2005).\n[9] S. Pirandola, D. Vitali, P. Tombesi, S. Lloyd, Phys. Rev. Lett.\n97, 150403 (2006).\n[10] D. Vitali, S. Mancini, and P. Tombesi, J. Phys. A: Math. Theor.\n40, 8055 (2007).7\n[11] M. J. Hartmann and M. B. Plenio, Phys. Rev. Lett. 101, 200503\n(2008).\n[12] S. Huang and G. S. Agarwal, New J. Phys. 11, 103044 (2009).\n[13] K. Borkje, A. Nunnenkamp, and S. M. Girvin, Phys. Rev. Lett.\n107, 123601 (2011).\n[14] M. Abdi, S. Pirandola, P. Tombesi, and D. Vitali, Phys. Rev.\nLett. 109, 143601 (2012).\n[15] Y .-D. Wang and A. A. Clerk, Phys. Rev. Lett. 110, 253601\n(2013).\n[16] H. Tan, G. Li, and P. Meystre, Phys. Rev. A 87, 033829 (2013).\n[17] H. Flayac and V . Savona, Phys. Rev. Lett. 113, 143603 (2014).\n[18] J.-Q. Liao, Q.-Q. Wu, and F. Nori, Phys. Rev. A 89, 014302\n(2014).\n[19] R.-X. Chen, L.-T. Shen, Z.-B. Yang, H.-Z. Wu, and S.-B.\nZheng, Phys. Rev. A 89, 023843 (2014).\n[20] M. J. Woolley and A. A. Clerk, Phys. Rev. A 89, 063805 (2014).\n[21] M. Abdi and M. J. Hartmann, New J. Phys. 17, 013056 (2015).\n[22] J. Li, I. Moaddel Haghighi, N. Malossi, S. Zippilli, and D. Vi-\ntali, New J. Phys. 17, 103037 (2015).\n[23] L. F. Buchmann and D. M. Stamper-Kurn, Phys. Rev. A 92,\n013851 (2015).\n[24] S. Zippilli, J. Li, and D. Vitali, Phys. Rev. A 92, 032319 (2015).\n[25] C. J. Yang, J. H. An, W. Yang, and Y . Li, Phys. Rev. A 92,\n062311 (2015).\n[26] O. Houhou, H. Aissaoui, and A. Ferraro, Phys. Rev. A 92,\n063843 (2015).\n[27] M. Asjad, S. Zippilli, and D. Vitali, Phys. Rev. A 93, 062307\n(2016).\n[28] M. Wang, X.-Y . L ¨u, Y .-D. Wang, J. Q. You, and Y . Wu, Phys.\nRev. A 94, 053807 (2016).\n[29] J. Li, G. Li, S. Zippilli, D. Vitali, and T.-C. Zhang, Phys. Rev.\nA95, 043819 (2017).\n[30] S. Kiesewetter, R. Y . Teh, P. D. Drummond, and M. D. Reid,\nPhys. Rev. Lett. 119, 023601 (2017).\n[31] S. Chakraborty and A. K. Sarma, Phys. Rev. A 97, 022336\n(2018).\n[32] H. Rudolph, K. Hornberger, and B. A. Stickler, Phys. Rev. A\n101, 011804(R) (2020).\n[33] A. K. Chauhan, O. Cernotik, and R. Filip, New J. Phys. 22,\n123021 (2020).\n[34] L. Martinetz, K. Hornberger, J. Millen, M. S. Kim, and B. A.\nStickler, npj Quantum Inf 6, 101 (2020).\n[35] G. Li and Z.-Q. Yin, arXiv:2111.11620.\n[36] J. Poyatos, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. 77, 4728\n(1996).\n[37] A. R. R. Carvalho, P. Milman, R. L. de Matos Filho, and L.\nDavidovich, Phys. Rev. Lett. 86, 4988 (2001).\n[38] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. B ¨uchler, and\nP. Zoller, Nature Phys. 4, 878 (2008).\n[39] F. Verstraete, M. M. Wolf, and J. I. Cirac, Nature Phys. 5, 633\n(2009).\n[40] S. Pielawa, L. Davidovich, D. Vitali, and G. Morigi, Phys. Rev.\nA81, 043802 (2010).\n[41] F. Fr ¨owis, P. Sekatski, W. D ¨ur, N. Gisin, and N. Sangouard,\nRev. Mod. Phys. 90, 025004 (2018).\n[42] A. Bassi, K. Lochan, S. Satin, T. P. Singh, and H. Ulbricht, Rev.\nMod. Phys. 85, 471 (2013).\n[43] S. Bose et al. , Phys. Rev. Lett. 119, 240401 (2017).\n[44] C. Marletto and V . Vedral, Phys. Rev. Lett. 119, 240402 (2017).\n[45] J. Li, S.-Y . Zhu, and G. S. Agarwal, Phys. Rev. Lett. 121,\n203601 (2018).\n[46] J. Li, S. Y . Zhu, and G. S. Agarwal, Phys. Rev. A 99, 021801(R)(2019).\n[47] J. Li and S.-Y . Zhu, New J. Phys. 21, 085001 (2019).\n[48] H. Tan, Phys. Rev. Research 1, 033161 (2019).\n[49] M.-S. Ding, L. Zheng, and C. Li, J. Opt. Soc. Am. B 37, 627\n(2020).\n[50] Z.-B. Yang, J.-S. Liu, A.-D. Zhu, H.-Y . Liu, and R.-C. Yang,\nAnn. Phys. 532, 2000196 (2020).\n[51] J. Li and S. Gr ¨oblacher, Quantum Sci. Technol. 6, 024005\n(2021).\n[52] M.-S. Ding, X.-X. Xin, S.-Y . Qin, and C. Li, Opt. Commun.\n490, 126903 (2021).\n[53] W. Zhang, D.-Y . Wang, C.-H. Bai, T. Wang, S. Zhang, and H.-F.\nWang, Opt. Express 29, 11773 (2021).\n[54] S.-F. Qi and J. Jing, Phys. Rev. A 103, 043704 (2021).\n[55] B. Sarma, T. Busch, and J. Twamley, New J. Phys. 23, 043041\n(2021).\n[56] Y .-T. Chen, L. Du, Y . Zhang, and J.-H. Wu, Phys. Rev. A 103,\n053712 (2021).\n[57] T.-X. Lu, H. Zhang, Q. Zhang, and H. Jing, Phys. Rev. A 103,\n063708 (2021).\n[58] W. Zhang, T. Wang, X. Han, S. Zhang, and H.-F. Wang, Opt.\nExpress 30, 10969 (2022).\n[59] C. Kittel, Phys. Rev. 110, 836 (1958).\n[60] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Sci. Adv. 2,\ne1501286 (2016).\n[61] C. A. Potts, E. Varga, V . Bittencourt, S. V . Kusminskiy, and J.\nP. Davis, Phys. Rev. X 11, 031053 (2021).\n[62] R.-C. Shen, J. Li, Z.-Y . Fan, Y .-P. Wang, and J. Q. You, Phys.\nRev. Lett. Phys. Rev. Lett. 129, 123601 (2022).\n[63] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Rev. Mod.\nPhys. 86, 1391 (2014).\n[64] M. Yu, H. Shen, and J. Li, Phys. Rev. Lett. 124, 213604 (2020).\n[65] J. Li, Y .-P. Wang, J. Q. You, and S.-Y . Zhu, Nat. Sci. Rev.\nnwac247 (2022).\n[66] D. Lachance-Quirion, Y . Tabuchi, A. Gloppe, K. Usami, and Y .\nNakamura, Appl. Phys. Express 12, 070101 (2019).\n[67] H. Y . Yuan, Y . Cao, A. Kamra, R. A. Duine, and P. Yan, Phys.\nRep. 965, 1 (2022).\n[68] J. Li, Y .-P. Wang, W.-J. Wu, S.-Y . Zhu, and J. Q. You. PRX\nQuantum 2, 040344 (2021).\n[69] F. Heyroth et al., Phys. Rev. Applied 12, 054031 (2019).\n[70] Z.-Y . Fan, H. Qian, and J. Li, Quantum Sci. Technol. 8, 015014\n(2023).\n[71] H. Huebl, C.W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein,\nA.Marx, R.Gross, and S. T. B.Goennenwein, Phys. Rev. Lett.\n111, 127003 (2013).\n[72] Y . Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and\nY . Nakamura, Phys. Rev. Lett. 113, 083603 (2014).\n[73] X. Zhang, C. L. Zou, L. Jiang, and H. X. Tang, Phys. Rev. Lett.\n113, 156401 (2014).\n[74] Z.-Y . Fan, R.-C. Shen, Y .-P. Wang, J. Li, and J. Q. You. Phys.\nRev. A 105, 033507 (2022).\n[75] C. Gonzalez-Ballestero, D. H ¨ummer, J. Gieseler, and O.\nRomero-Isart, Phys. Rev. B 101, 125404 (2020).\n[76] R. Benguria and M. Kac, Phys. Rev. Lett. 46, 1 (1981); V . Gio-\nvannetti and D. Vitali, Phys. Rev. A 63, 023812 (2001).\n[77] G. Vidal and R. F. Werner, Phys. Rev. A 65, 032314 (2002); M.\nB. Plenio, Phys. Rev. Lett. 95, 090503 (2005).\n[78] T. A. Palomaki, J. D. Teufel, R. W. Simmonds, and K. W. Lehn-\nert, Science 342, 710 (2013)." }, { "title": "1202.6166v1.Gallium_Substituted__114__YBaFe4O7__From_a_ferrimagnetic_cluster_glass_to_a_cationic_disordered_spin_glass.pdf", "content": " 1 Gallium Substituted “114” YBaFe 4O7: From a ferrimagnetic cluster glass to \na cationic disordered spin glass \n \nTapati Sarkar *, V. Caignaert, V. Pralong and B. Raveau \n \n \nLaboratoire CRISMAT, UMR 6508 CNRS ENSICAEN, \n6 bd Maréchal Juin, 14050 CAEN, France \n \nAbstra ct \n \n The study of the ferrites YBaFe 4-xGaxO7 shows that the substitution of Ga for Fe in \nYBaFe 4O7 stabilizes the hexagonal symmetry for 0.40 \n x \n 0.70, at the expense of the cubic \none. Using combined measurements of a. c. and d. c. magnetization, we estab lish that Ga \nsubstitution for Fe in YBaFe 4O7 leads to an evolution from a geometrically frustrated spin \nglass (for x = 0) to a cationic disorder induced spin glass (x = 0.7 0). We also find an \nintermediate narrow range of doping where the samples are clearl y phase separated having \nsmall ferrimagnetic clusters embedded in a spin glass matrix. The origin of the ferrimagnet ic \nclusters lies in the change in symmetry of the samples from cubic to hexagonal (and a \nconsequent lifting of the geometric al frustration) as a result of Ga doping. We also show the \npresence of exchange bias and domain wall pinning in these samples. The cause of both these \neffects can be traced back to the inherent ph ase separation present in the samples. \n \n \n \n \n \n \n \n \nKeywords : “114” oxides, magne tic frustration , phase separation . \n \n \n \n \n* Corresponding author: Dr. Tapati Sarkar \ne-mail: tapati.sarkar @ensicaen.fr \nFax: +33 2 31 95 16 00 \nTel: +33 2 31 45 26 32 2 Introduction \n \n The recent studies of the “114” cobaltites (Ln,Ca) 1BaCo 4O7 [1 – 5] and ferrit es \n(Ln,Ca) 1BaFe 4O7 [6 – 8] have generated a lot of interest in the scientific community because \nof their complex magnetic, electronic and thermoelectric properties [9]. These cobaltites and \nferrites have the same basic structure, and are closely related to spinels and barium \nhexaferrites by their close packing of “O 4” and “BaO 3” layers. This close packing forms a 3 -\ndimensional framework [Fe 4O7]\n (or [Co 4O7]\n) consisting of corner -sharing FeO 4 (or CoO 4) \ntetrahedra , with the lanthanide elements occupying the octahedral sites of this framework. The \ntriangular geometry of the cobalt (or iron) sublattices ( Fig. 1 ) plays a dominant role in their \nmagnetic properties. It was indeed shown that for hexagonal LnBaCo 4O7 cobaltites ( Fig. 1 \n(a)), there exists a strong com petition between the 1 D magnetic ordering along the \n \ndirection in the “Co 5” trigonal bipyramids, and the magnetic frustration in the (001) plane \nbuilt up of “Co 3” triangles [10, 11]. In fact, the magnetic frustration can be lifted by an \northorhombic dist ortion of the structure. This is illustrated by the concomitant structural and \nmagnetic transitions that appear at low temperature in these cobaltites [1, 2, 12], and by the \nferrimagnetic structure of CaBaCo 4O7 [13]. Similarly, the “114” ferrites exhibit a competition \nbetween 1 D magnetic ordering and 2 D magnetic frustration, as has been shown for the \nhexagonal phases CaBaFe 4O7 [6], and for CaBa Fe4-xLixO7 [14]. But importantly, the “114” \nferrites differ from the “114” cobaltites by the fact that the LnBaFe 4O7 oxides exhibit a cubic \nstructure [7]. Though the latter is closely related to the hexagonal structure, the iron sublattice \nis very different ( Fig. 1 (b) ), consisting of “Fe 4” tetrahedra instead of “Fe 5” bipyramids and \n“Fe 3” triangles. No structural tra nsition appears at low temperature, and consequently, the \ncubic ferrites exhibit a spin glass behaviour due to a perfect geometrical frustration. Further, \nthe LnBaFe 4O7 series exhibit s an oxidation state disorder. Unlike the case of CaBaCo 4O7 [13], \nno char ge ordering is observed in LnBaFe 4O7, and this disorder is also important for the \nobserved glassiness. \nRecently, we showed that the substitution of a divalent cation, Zn2+, for iron in \nYBaFe 4O7, allowed the hexagonal symmetry to be stabilized at the detrim ent of the cubic one \n[15]. Paradoxically, it was observed that the substitution of this diamagnetic cation for Fe2+ \ninduces ferrimagnetism, in contrast to the spin glass behaviour of the undoped phase \nYBaFe 4O7. In fact, a competition between ferrimagnetism and magnetic frustration was \nobserved for the hexagonal phase YBaFe 4-xZnxO7. This was interpreted as the effect of two 3 antagonist phenomena: the partial lifting of the geometrical frustration due to the appearance of \nthe hexagonal symmetry inducing a 1 D magnetic ordering, and the existence of cationic \ndisordering favouring the glassy state. \n Bearing in mind that the Fe2+:Fe3+ ratio is a crucial factor governing the magnetic \nproperties of iron oxides, it must be emphasized that the substitution of Zn2+ for Fe2+ increases \nthe average valence of iron , i.e. the Fe2+:Fe3+ ratio decreases from 3 in the spin glass phase \nYBaFe 4O7 to 2.6 – 1.5 in the solid solution YBaFe 4-xZnxO7 when x changes from 0.4 to 1.5 \n[15]. In order to further understand the role of the ave rage valence of iron in the magnetic \nproperties of these ferrites, we have investigated the possibility of substitution of a diamagnetic \ncation such as gallium for Fe3+ in the YBaFe 4O7 structure. In the present study of the ferrite \nYBaFe 4-xGaxO7, we show t hat the introduction of gallium in the structure stabilizes the \nhexagonal symmetry, similar to the zinc substitution, but differently from the latter, the lifting \nof the geometrical frustration induces the formation of ferrimagnetic clusters embedded in a \nspin glass matrix, which tend to disappear as the gallium content increases, leading to a pure \nspin glass for higher Ga content, with a higher T g compared to YBaFe 4O7. \n \nExperimental \n \nPhase -pure samples of YBaFe 4-xGaxO7 [x = 0.4 0 – 0.70] were prepared by solid state \nreaction technique. The precursors used were Y 2O3, BaFe 2O4, Ga 2O3, Fe 2O3 and metallic Fe \npowder. First, the precursor BaFe 2O4 was prepared from a stoichiometric mixture of BaCO 3 \nand Fe 2O3 annealed at 1200°C for 12 hrs in air. In a second step, a stoichiometric mixture of \nY2O3, BaFe 2O4, Ga 2O3, Fe 2O3 and metallic Fe powder was intimately ground and pressed in \nthe form of rectangular bars. The bars were then kept in an alumina finger, sealed in silica \ntubes under vacuum and annealed at 1100°C for 1 2 hrs. Finally, the samples were quenched to \nroom temperature in order to stabilize the “114” phase. \nThe X -ray diffraction patterns were registered with a Panalytical X’Pert Pro \ndiffractometer under a continuous scanning mode in the 2\n range 10° - 120° an d step size \n 2\n \n= 0.017°. The cationic composition was confirmed by means of Energy Dispersive X -Ray \nSpectroscopy (EDS) technique using a Scanning Electron Microscope (ZEISS Supra 55). The \nd. c. magnetization measurements were performed using a superconduc ting quantum \ninterference device (SQUID) magnetometer with variable temperature cryostat (Quantum \nDesign, San Diego, USA). The a. c. susceptibility, \n ac(T) was measured with a PPMS from \nQuantum Design with the frequency ranging from 10 Hz to 10 kHz. H ac was kept fixed at 10 4 Oe, while H dc was varied from 0 Oe to 2000 Oe. All the magnetic properties were registered on \ndense ceramic bars of dimensions ~ 4 \n 2 \n 2 mm3. \n \nResults and discussion \n \n \n Similar to Zn substitution, Ga substitution also favours the forma tion of the hexagonal \nphase at the expense of the cubic one. Nevertheless, the homogeneity range of the hexagonal \nYBaFe 4-xGaxO7 solid solu tion is significantly different (0.40 \n x \n 0.70) vis – à – vis that of \nYBaFe 4-xZnxO7 [15]. The cubic symmetry of YBaF e4O7 is retained for 0 \n x \n 0.20, whereas \nthe domain 0.20 < x < 0.40 is biphasic, corresponding to a mixture of the cubic and hexagonal \nphases. On the other hand, for x > 0.70 , several impurity phases appear , namely Y 2O3 and \nGa2O3. The cationic compositio n of the single phase obtained for the range 0.40 \n x \n 0.70 \nusing EDS analysis are also shown in Table 1 . \n \nStructural characterization \n \n In Fig. 2, we show the X -ray diffraction (XRD) pattern s of the two end members, (a) \nYBaFe 3.6Ga0.4O7 and (b) YBaFe 3.3Ga0.7O7 as representative example s. As stated before, t he \nsample s are seen to stabilize in the hexagonal symmetry with the space group P63mc. The \nRietveld analysis of the lattice structure was done using the FULLPROF refinement program \n[16] and the fit s are also shown in Fig. 2. All the samples in the range x = 0.4 0 – 0.70 were \nseen to stabilize in the same hexagonal symmetry. \nThe extracted cell parameters have been tabulated in Table 1 . The ionic radius of Fe3+ (0.49 \nÅ) is very similar to that of Ga3+ (0.47 Å). As can be seen from the extracted cell parameters \nshown in Table 1, a increases very slightly as x increases (an increase of only ~ 0.08 % as x \nincreases from 0.4 to 0.7), while c shows a slight decrease (~ 0.11 %). This causes the cell \nvolume to re main practically unchanged as a function of doping in accordance with the \nsimilar ionic radii of Fe3+ and Ga3+. \n \nD. C. magnetization studies \n \n In the “114” ferrites, it has been established earlier [8, 15] that ferrimagnetism is \ninherently linked with the cross -over from cubic to hexagonal symmetry. The doping -induced \ntransition to the hexagonal symmetry involves a partial lifting of the 3D geometrical 5 frustration, which is the root cause of the appearance of ferrimagnetism. Thus, we restrict our \ndiscussion of the magnetic data to the YBaFe 4-xGaxO7 samples exhibiting hexagonal symmetry \n(0.4 \n x \n 0.7). We note here that the cubic samples (x < 0.2) are spin glasses similar to the \nundoped YBaFe 4O7, and will not be discussed further. \n The temperature dependence of d. c. magneti c susceptibility (\n dc = M/H) was registered \naccording to the standard zero field cooled (ZFC) and field cooled (FC) procedures. A \nmagnetic field of 0.3 T was applied during the measurements. The measurements were done in \na temperature rang e of 5 K to 300 K. The \n ZFC(T) and \n FC(T) curves of all the samples are \nshown in Fig. 3. \n Undoped YBaFe 4O7 is a spin glass with T g = 50 K [ 7]. The ZFC \n dc versus T curve for \nYBaFe 4O7 shows a pure cusp -like shape [ 7], typical of canonical spin glasses. A close look at \nFig. 3 reveals that the YBaFe 4-xGaxO7 series of samples shows two different kinds of low \ntemperature M ZFC(T) curves vis -à-vis the shape of the curves. While the \n dc(T) curve of the \nhighest substituted sample (x = 0.7 0) is very similar to that o f canonical spin glasses (with a \npeak at ~ 50 K and a gradual decrease of the magnetization value below 50 K), for the lowest \ndoped sample (x = 0.4 0), there is a sharp drop in the susceptibility value below the temperature \nat which \n ZFC reaches its maximum value (75 K). The susceptibility value drops sharply till ~ \n50 K (marked by a black arrow in Fig. 3 (a)), below which the decrease in \n ZFC is more \ngradual. We note here that the measuring field that we have chosen (0.3 T) is smaller than the \ncoercive fiel d of the YBaFe 3.6Ga0.4O7 sample at T = 5 K (data shown later in Fig. 5 ). Thus, it is \nquite possible that the sharp drop in the susceptibility value occurs at the temperature where the \ncoercive field of the sample becomes smaller than 0.3 T. However, follow ing this argument, \nwe should have obtained similar sharp drops in the \n ZFC(T) curves for the x = 0.5 and x = 0.6 \nsamples also, as the coercive fields of the x = 0.5 and x = 0.6 samples at T = 5 K are also larger \nthan 0.3 T. Instead , it is observed that the sharp drop in the \n ZFC(T) curve seen in the x = 0.40 \nsample is reduced to small kinks in the \n ZFC(T) curves for the x = 0.50 and x = 0.60 samples. \nMore importantly, a study of the temperature dependence of the coercive field (H C) of the x = \n0.4 sample (d ata not shown here) reveals that H C becomes smaller than 0.3 T at ~ 17 K (i.e. at \na temperature much below 50 K). This suggests that this feature is not a simple effect of the \ncoercive field, rather it may have a more complex origin. \nAnother possibility is that this sudden decrease in \n ZFC(T) is due to domain wall \npinning effects, which has, in fact, been observed previously in manganites [1 7 – 19]. Due to \npinning, the domains would not freely rotate below the pinning temperature unless a high 6 enough extern al field is present to overcome the pinned state. Upon zero field cooling, the \ndomains would be pinned into random orientations. Whe n a low field is applied (0.3 T in this \ncase), the pinning effect still dominates over the effect of the applied magnetic fi eld, and the \nmagnetization is lower than what would be expected in the absence of pinning. However, the \npinned domain walls can be thermally activated by increasing the temperature. This could be \nthe cause of the visible jump in the \n ZFC(T) curve at the te mperature where the pinning effects \nare overcome by temperature (~ 50 K). As can be seen in Fig. 3, the pinning effect gradually \ndecreases as the doping concentration is increased (the sharp drop in the \n ZFC(T) curve seen in \nthe x = 0.4 0 sample is reduced to small kinks in the \n ZFC(T) curves for the x = 0.5 0 and x = \n0.60 samples, and completely vanishes for the x = 0.7 0 sample). This indicates that the domain \nwall pinning is more prominent for small doping and vanishes for higher doping. This is \ncounter -intuitive if the pinning is thought to arise due to the presence of Ga in the lattice. Thus, \nthe fact that the domain wall pinning decreases with an increase in the doping concentration \nleads us to believe that this pinning does not arise from the disorder in the system. Rather, it \nhas a more complex origin, which we discuss later. \n Before we proceed further, we perform some additional measurements to make sure \nthat the sharp drop in the \n ZFC(T) curve observed in the lowest doped sample is indeed due to \ndomain wall effects , and not arising from some additional (antiferro) magnetic transition in the \nsample. Thus, we subject the x = 0.4 0 sample to a degaussing experiment [1 7], wherein the \nsample was initially cooled from 300 K down to 5 K in a zero external magne tic field. At 5 K, \na large magnetic field (5 T) was applied. The magnetic field was then reduced to zero, and the \nsample was degaussed at 5 K by cycling a field of reducing intensity so that the remanent \nmagnetization of the sample was reduced to zero. A m agnetic field of 0.3 T was then applied, \nand the \n dc(T) curve was recorded while warming the sample, in the same wa y as ZFC \nmagnetization is recorded. The results are shown in Fig. 4. We find that the sudden sharp drop \nobserved in the normal ly obtained ZFC curve ( Fig. 4 (a)) vanishes when the sample is \nsubjected to a high enough magnetic field, and then degaussed (Fig. 4 (b)). This experiment, \nthus, provides supplementary evidence that the sudden drop in magnetization seen below 75 K \nis not due to any kind of (antiferro) magnetic transition in the sample , but is probably \nassociated with domain wall pinning effects. The high magnetic field (5 T) to which the sample \nwas subjected was sufficient for domain wall displacements thereby destroying the pinning. In \nfact, a ZFC magnetization recorded under a high enough field of 5 T (see inset of Fig. 4 (a)) \ndoes not show any sharp drop in the magnetization of the sample. 7 The d. c. magnetization M(H) curves of all the samples registered at T = 5 K are shown \nin Fig. 5. The virgin curves of the M(H) loops are represented by black circles while the rest of \nthe M(H) loops are shown by red lines. The first notable point is that for higher Ga substitution \n(x = 0.7 0), the M(H) loop is narrow and S – shaped ( Fig. 5 (d)), which is quite typical of spin \nglasses and superparamagnets . On the other hand, for lower Ga substitution, the samples have \nlarger loops with higher values of the c oercivity and remanent magnetization, which keep \nincreasing as the doping concentration is decreas ed. This indicates the presence of a higher \ndegree of magnetic ordering in the lower doped samples as compared to the higher doped ones. \n Another feature which strongly supports the presence of domain wall pinning is that the \nvirgin curve of the x = 0.4 0 sample lies slightly outside the main M(H) loop ( Fig. 5 (a)). This \nunusual feature of the virgin curve lying outside the main hysteresis loop has earlier been \nassociated with irreversible domain wall motion in spinel oxides [ 20]. We also note that the \nvirgi n curve starts to shift inside the main M(H) loop as the doping concentration (x) is \nincreased, and for the x = 0.7 0 sample, the entire virgin curve lies inside the main M(H) loop \n(Fig. 5 (d)). We again note that the domain wall pinning is more prominent i n the samples with \nlower doping concentration. \n In Fig. 6, we once again show the d. c. magnetization M(H) curves of all the samples \nregistered at T = 5 K, but in three different modes: (i) normal ZFC mode, (ii) FC mode with a \nmagnetic field of 2 T and (iii) FC mode with a magnetic field of H = - 2 T. In the ZFC mode, \nthe samples were cooled from 300 K to 5 K in zero external magnetic field, following which M \nversus H curves were registered. In the FC mode, on the other hand, the samples were cooled \nfrom 30 0 K to 5 K in the presence of an external magnetic field (H = 2 T or – 2 T), and then M \nversus H curves were registered. For the highest substituted sample (x = 0.7 0), all three M(H) \ncurves overlap each other ( Fig. 6 (d)). However, for lower doping concent ration, the field \ncooled M(H) loops exhibit shifts both in the field as well as in the magnetization axes. This is \nthe exchange bias phenomenon [ 21, 22] that results from exchange interaction between \nferromagnetic and antiferromagnetic materials. In our YB aFe 4-xGaxO7 samples with low Ga \nconcentration, the observed exchange bias can be explained in terms of interfacial exchange \ncoupling between the coexisting ferrimagnetic cluster glass and the disordered spin glass -like \nphases. This exchange bias effect ari sing from the inherent phase separation in the YBaFe 4-\nxGaxO7 samples is similar to that seen in some disordered manganites [ 23]. As can be seen \nfrom Fig. 6, the exchange bias effect keeps decreasing as the doping concentration is increased, \nand as stated b efore, it completely disappears for the doping concentration x = 0 .70. We can \nexplain this observation by considering that for lower Ga concentration, the samples consist of 8 coexisting ferrimagnetic clusters embedded in a spin glass -like matrix, but as the doping \nconcentration by the diamagnetic cation (Ga) is increased, the ferrimagnetic clusters are \nprogressively reduced and we ultimately get a homogeneous spin glass (x = 0.7 0). The absence \nof phase separation in the x = 0.7 0 sample, thus, results in an a bsence of the exchange bias \neffect. The fact that the YBaFe 4-xGaxO7 samples with lower Ga concentration are intrinsically \nphase separated, while the x = 0.7 0 sample is not, also affords us an alternative explanation for \nthe domain wall pinning effects seen in the lower doped samples. As stated previously, the fact \nthat the domain wall pinning is seen in the lower doped samples and not in th e x = 0.7 0 sample \nmeans that it cannot arise from the disorder in the system. Rather, we believe that the pinning \narise s from an interplay between the two magnetic phases in the phase separated samples. Such \na domain wall pinning process arising from the interplay between two coexisting magnetic \nphases has been seen earlier in intermetallic alloys [ 24]. We also note that a part from the \nexchange bias effect in the lower doped samples, field cooling also results in an overall \nincrease in the coercivity and remanence magnetization values. This can be interpreted as an \nincrease in the volume fraction of the magnetically ordered phase when the samples are cooled \nin the presence of an external magnetic field. Since field cooling improves the remanence in \nthe lower doped samples, hence it was important to register M -H curves after field cooling \nwith positive as well as negative coo ling fields and check whether the M -H loops shift in \nopposite directions in order to confirm that there is indeed a genuine exchange bias effect in \nthe lower doped samples. \n \nA. C. magnetic susceptibility studies \n \n The measurements of the a. c. magnetic sus ceptibility \n ac(T, f, H) were performed at \ndifferent frequencies ranging from 10 Hz to 10 kHz, and different external magnetic field s \n(Hdc) ranging from 0 T to 0.2 T using a PPMS facility. The amplitude of the a. c. magnetic \nfield was ~ 0.001 T In Fig. 7 and Fig. 8, we show the temperature dependence of the real (in -\nphase) component of the a. c. susceptibility in the temperature range 10 K – 160 K of the \nlowest doped sample (x = 0.4 0) and the highest doped sample (x = 0.7 0) respectively, with a \nmeasuring f requency of 10 kHz and in zero magnetic field ( Hdc = 0). \nFrom Fig. 7, it is clear that the \n '(T) curve of the x = 0.4 0 sample shows two features, \none at T = 86 K (marked by a black arrow), and the second at T = 44 K (marked by a red \narrow). While the high temperature feature at 86 K is a clear peak, the low temperature one at \n44 K is a shoulder like feature and is more clearly evidenced in the imaginary (out -of-phase) 9 component of the a. c. susceptibity (shown in inset (a) of Fig. 7). Repeating the measure ments \nusing four measuring frequencies (ranging from 10 Hz – 10 kHz) reveals that both the features \nare frequency dependent (this is shown in inset (b) of Fig. 7). The x = 0.7 0 sample, on the other \nhand, shows only one feature (at T = 60 K) in the \n '(T) an d \n''(T) curves, shown in the main \npanel and inset (a) of Fig. 8 respectively. Inset (b) of Fig. 8 shows the \n '(T) curves measured \nusing four different frequencies, and reveals that this peak at 60 K is also strongly frequency \ndependent. We note that the i maginary part of \n for YBaFe 3.6Ga0.4O7 is ~ 8 % of the real part \nof \n . This is commonly found for systems where the spin domains are relatively large. \nYBaFe 3.6Ga0.4O7 can thus be described as a cluster glass. On the other hand, the imaginary part \nof \n for YBaFe 3.3Ga0.7O7 is significantly smaller (about 2.4% of the real part of \n ) which \nindicates that YBaFe 3.3Ga0.7O7 is closer to a canonical spin glass. \n Although all the features in the a. c. susceptibility curves of the two samples described \nabove have a si ngle commonality , in that they are all strongly frequency dependent, but the \nnature a nd origin of these peaks can be quite different. Specifically, we need to establish the \nnature of the low temperature shoulder in the x = 0.4 sample at T ~ 50 K. Since it occurs close \nto the temperature where we ha ve evidenced domain wall pinning from the d. c. magnetic data, \nit is tempting to attribute this low temperature shoulder in the a. c. susceptibility data to the \nsame phenomenon. However, it is also possible that t his low temperature feature is a \nsuperparamagnetic effect of the ferrimagnetic clusters. To investigate this , we perform further \nmeasurements of \n '(T) and \n ''(T) of the two limiting samples (x = 0.4 0 and x = 0.7 0) in the \npresence of different external magn etic fields H dc ranging from 0 to 0.2 T. The results for the \nYBaFe 3.3Ga0.7O7 sample are shown in Fig. 9. It is seen that both \n ' and \n '' are strongly \nsuppressed by the magnetic field. The peak also shows a continual shift towards lower \ntemperature as the e xternal magnetic field is increased (see the black arrows in Fig. 9). This is \ntypical of the behaviour of a spin glass freezing temperature under the influence of magnetic \nfield. \n In Fig. 10, we show the results for the YBaFe 3.6Ga0.4O7 sample. It is seen t hat w hile the \nhigh temperature peak was significantly suppressed in the presence of external magnetic field, \nthe low temperature peak was largely unaffected relative to the case H dc = 0. This rules out the \nscenario of superparamagnetism being responsible f or this low temperature feature, and \nconfirms that the 50 K anomaly arises due to enhanced domain wall pinning, signatures of \nwhich have been observed and commented upon earlier in the d. c. magnetization \nmeasurements also. We also note that the high tempe rature peak shifts towards higher 10 temperature as the external magnetic field is increased (see the black arrow in Fig. 10 (a)). \nThis is not expected for a pure spin glass freezing transition, where the peak should shift \ntowards lower temperature as the mag netic field is increased. We believe that this anomaly \narises because the x = 0.4 0 sample is not a pure spin glass, rather it is a phase separated sample \nconsisting of ferrimagnetic clusters embedded in a spin glass matrix. \n \nConclusion \n \n These results show that the substitution of Ga3+ for Fe3+ in YBaFe 4O7 induces a \nstructural transition from cubic to hexagonal, similar to the substitution of Zn2+ for Fe2+ in this \ncompound. Though the two types of substitutions induce a lifting of the geometrical frustratio n \nthrough a change of the structure, the effect of these diamagnetic cations upon the magnetic \nproperties is different. A strong ferrimagnetic component is induced by zinc substitution [15], \nwhereas Ga substitution leads to the formation of ferrimagnetic c lusters embedded in a spin \nglass matrix, essentially leading to phase separation in the samples. The difference originates \nfrom the opposite evolution of the Fe3+:Fe2+ ratio as the substitution rate increases in the two \ncases. Both Fe3+ and Fe2+ exhibit th e high spin configuration since they have a tetrahedral \ncoordination in these ferrites. Thus, the magnetic moment induced by the eg2t2g3 Fe3+ cations \nshould be much higher than that induced by the eg3t2g3 Fe2+ cations. Hence, an increase in the \nFe3+:Fe2+ ratio should favour stronger magnetic interactions. In the case of Zn2+ doping, the \nFe3+:Fe2+ ratio increases , thereby favouring the appearance of ferrimagnetism. On the other \nhand, for Ga3+ doping, the Fe3+:Fe2+ ratio decreases , thereby inducing only weak \nferrimagnetism and cluster formation . In both series, YBaFe 4-xGaxO7 and YBaFe 4-xZnxO7, a \ndilution effect is observed with an increase in the doping concentration. As a consequence, \nferrimagnetism is weakened for higher concentrations in the Zn – phase. In the Ga – phase, the \nferrimagnetic clusters are magnetically coupled by exchange interactions mediated through the \nsurrounding spin glass matrix. For higher Ga concentrations, the exchange coupling between \nthe ferrimagnetic clusters becomes less efficient, ultimately leading to the formation of a pure \nspin glass phase for x = 0.70, which is similar to the pristine sample (x = 0), but with a slightly \nhigher T g (60 K). The Ga – substituted phase also differs from the Zn – phase by the presence \nof exchange bias and domain wall pinning. The cause of both these effects can be traced back \nto the inherent phase separation present in the samples. \n \n 11 Acknowledgements \n \nWe acknowledge the CNRS and the Conseil Regional of Basse Normandie for financial \nsupport in the frame of Emergence Program and N°10P01391 . V. P. acknowledges support by \nthe ANR -09-JCJC -0017 -01 (Ref: JC09_442369). \n \nReferences : \n \n [1] Martin Valldor and Magnus Andersson, Solid State Sciences , 2002 , 4, 923 \n [2] Martin Valldor, J. Phys.: Con dens. Matter ., 2004 , 16, 9209 \n [3] A. Huq, J. F. Mitchell, H. Zheng, L. C. Chapon, P. G. Radaelli, K. S. Knight and P. \n W. Stephens, J. Solid State Chem ., 2006 , 179, 1136 \n [4] D. D. Khalyavin, L. C. Chapon, P. G. Radaelli, H. Zheng and J. F. Mitchell, Phys. \n Rev. B , 2009 , 80, 144107 \n [5] V. Caignaert, V. Pralong, A. Maignan and B. Raveau, Solid State Communications , \n 2009 , 149, 453 \n [6] B. Raveau, V. Caignaert, V. Pralong, D. Pelloquin and A. Maignan, Chem. Mater ., \n 2008 , 20, 6295 \n [7] V. Caignaert, A. M. Abakumov, D. Pelloquin, V. Pralong, A. Maignan, G. Van \n Tendeloo and B. Raveau, Chem. Mater ., 2009 , 21, 1116 \n [8] V. Pralong , V. Caignaert, A. Maignan and B. Raveau, J. Mater. Chem ., 2009 , 19, \n 8335 \n [9] B. Raveau, V. Caignaert, V. Pralong and A. Maignan, Z. Anorg. Allg. Chem ., 2009 , \n 635, 1869 \n [10] L. C. Chapon, P. G. Radaell i, H. Zheng and J. F. Mitchell, Phys. Rev. B , 2006 , 74, \n 172401 \n [11] P. Manuel, L. C. Chapon, P. G. Radaelli, H. Zheng and J. F. Mitchell, Phys. Rev. \n Lett., 2009 , 103, 037202 \n [12] A. Maignan, V. Caignaert, D. Pelloquin, S. Hébert, V. Pralong, J. Hejtmanek and D. \n Khomskii, Phys. Rev. B , 2006 , 74, 165110 \n [13] V. Caignaert, V. Pralong, V. Hardy, C. Ritter and B. Raveau, Phys. Rev. B , 2010 , 81, \n 094417 \n [14] K. Vijay anandhini, Ch. Simon, V. Pralong, V. Caignaert and B. Raveau, Phys. Rev. \n B, 2009 , 79, 224407 12 [15] T. Sarkar, V. Pralong , V. Caignaert and B. Raveau , Chem. Mater ., 2010, 22, 2885 \n [16] J. Rodriguez -Carvajal, An Introduction t o the Program FULLPROF 2000; Laboratoire \n Léon Brillouin, CEA -CNRS: Saclay, France (2001) \n [17] P. A. Joy and S. K. Date, J. Magn. Magn. Mater ., 2000 , 220, 106 \n [18] C. R. Sankar and P. A. Joy, Phys. Rev. B , 2005 , 72, 024405 \n [19] T. Gao, S. X. Cao, K. Liu, B. J. Kang, L. M. Yu, S. J. Yuan and J. C. Zhang, Journal \n of Phys: Conf. Series , 2009 , 150, 042038 \n [20] P. A. Joy and S. K. Date, J. Magn. Magn. Mater ., 2000 , 210, 31 \n [21] R. L. Stamps, J. Phys. D, 2000 , 33, R247 \n [22] W. H. Meiklejohn and C. P. Bean , Phys. Rev ., 1956 , 102, 1413 \n [23] S. Karmakar, S. Taran, E. Bose, B. K. Chaudhuri, C. P. Sun, C. L. Huang and H. D. \n Yang, Phys. Rev. B , 2008 , 77, 144409 \n [24] A. Bracchi, K. Samwer, S. Schneider and J. F. Löffler, Appl. Phys. Lett ., 2003 , 82, \n 721 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n 13 Table caption s \n \nTable 1 : Cell parameters as obtained from the Rietveld refinement of X -ray powder diffraction \ndata. \n \nFigure captions \n \nFigure 1: Schematic representation of (a) hexagonal LnBaCo 4O7 and (b) cubic YBaFe 4O7 \n(adapted from Ref. 8). For details, see text. \n \nFigure 2: X-ray diffraction pattern along with the fits for (a) YBaFe 3.6Ga0.4O7 and (b) \nYBaFe 3.3Ga0.7O7. \n \nFigure 3: Temperature dependence of the magnetic susceptibility (\n dc = M/H) collected \naccording to zero field cooling (ZFC) and field cooling (FC) processes for YBaFe 4-xGaxO7 (a) \nx = 0.40, (b) x = 0.50, (c) x = 0.60 and (d) x = 0.70, measured at B = 0.3 T. \n \nFigur e 4: \nZFC(T) curves of YBaFe 3.6Ga0.4O7 recorded (a) in the ZFC mode without \ndegaussing, and (b) after applying a magnetic field of 5 T and degaussing the ZFC sample (see \ntext for details). The inset in (a) shows \n ZFC(T) recorded under a magnetizing field o f 5 T. \n \nFigure 5: M (H) curves for YBaFe 4-xGaxO7 (a) x = 0.40, (b) x = 0.50, (c) x = 0.60 and (d) x = \n0.70, registered at T = 5 K . The virgin curves are shown in black circles, while the rest of the \nhysteresis loops are shown in red lines. \n \nFigure 6: M (H) curves for YBaFe 4-xGaxO7 (a) x = 0.40, (b) x = 0.50, (c) x = 0.60 and (d) x = \n0.70, registered at T = 5 K, measured after zero field cooling (red open circles), field cooling in \na field of 2 T (black lines) and field cooling in a field of - 2 T (blue line s). \n \nFigure 7 : Temperature dependence of the real (in -phase) component of a. c. susceptibility for \nYBaFe 3.6Ga0.4O7 as a function of temperature measured in zero magnetic field (H dc = 0), using \na frequency of 10 kHz. Inset (a) shows the imaginary (out -of-phase) component of the a. c. \nsusceptibility, and inset (b) shows the real (in -phase) component of the a. c. susceptibility \nmeasured using four different frequencies. 14 \nFigure 8 : Temperature dependence of the real (in -phase) component of a. c. susceptibility for \nYBaFe 3.3Ga0.7O7 as a function of temperature measured in zero magnetic field (H dc = 0), using \na frequency of 10 kHz. Inset (a) shows the imaginary (out -of-phase) component of the a. c. \nsusceptibility, and inset (b) shows the real (in -phase) component of the a. c. susceptibility \nmeasured using four different frequencies. \n \nFigure 9 : The (a) real (in -phase) and (b) imaginary (out -of-phase) component of a. c. \nsusceptibility for YBaFe 3.3Ga0.7O7 as a function of temperature. The driving frequency was \nfixed a t f = 1 kHz and H ac = 10 Oe. Each curve was obtained under different applied static \nmagnetic field (H dc) ranging from 0 T to 0.2 T. \n \nFigure 10 : The (a) real (in -phase) and (b) imaginary (out -of-phase) component of a. c. \nsusceptibility for YBaFe 3.6Ga0.4O7 as a function of temperature. The driving frequency was \nfixed at f = 1 kHz and H ac = 10 Oe. Each curve was obtained under different applied static \nmagnetic field (H dc) ranging from 0 T to 0.2 T. \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n 15 Table 1 \n \nDoping \nconcentration \n(x) \nCrystal \nsystem \n(Space \ngroup) \n \nUnit cell parameters \n c / a \n2 x as \nobtained \nfrom EDS \nanalysis a (Å) c (Å) \n0.40 Hexagonal \n(P63mc) \n6.320 (1) \n 10.383 (1) 1.6428 3.02 0.43 (2) \n0.50 Hexagonal \n(P63mc) \n6.322 (1) \n 10.376 (1) 1.6413 3.15 0.50 (1) \n0.60 Hexagonal \n(P63mc) \n6.323 (1) \n 10.37 4 (1) 1.6407 2.94 0.59 (1) \n0.70 Hexagonal \n(P63mc) \n6.325 (1) \n 10.37 2 (1) 1.6398 3.46 0.74 (3) \n \n \n \n \n \n \n \n \n \n \n \n \n 16 \n \n \nFig. 1 . Schematic representation of (a) hexagonal LnBaCo 4O7 and (b) cubic YBaFe 4O7 \n(adapted from Ref. 8). For details, s ee text. \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n 17 \n \n \nFig. 2. X-ray diffraction pattern along with the fit s for (a) YBaFe 3.6Ga0.4O7 and (b) \nYBaFe 3.3Ga0.7O7. \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n 18 \n \n \nFig. 3. Temperature dependence of the magnetic susceptibility (\n dc = M/H) collected according \nto zer o field cooling (ZFC) and field cooling (FC) processes for YBaFe 4-xGaxO7 (a) x = 0.40, \n(b) x = 0.50, (c) x = 0.60 and (d) x = 0.70, measured at H = 0.3 T. \n \n \n \n \n \n \n \n \n \n \n \n \n \n 19 \n \n \nFig. 4. \nZFC(T) curves of YBaFe 3.6Ga0.4O7 recorded (a) in the ZFC mode without dega ussing , \nand (b) after applying a magnetic field of 5 T and degaussing the ZFC sample (see text for \ndetails). The inset in (a) shows \n ZFC(T) recorded under a magnetizing field of 5 T. \n \n \n \n \n \n \n \n \n \n \n \n \n 20 \n \n \nFig. 5. M (H) curves for YBaFe 4-xGaxO7 (a) x = 0.40, (b) x = 0.50, (c) x = 0.60 and (d) x = \n0.70, registered at T = 5 K . The virgin curves are shown in black circles, while the rest of the \nhysteresis loops are shown in red lines. \n \n \n \n \n \n \n \n \n \n \n \n \n \n 21 \n \n \nFig. 6. M (H) curves for YBaFe 4-xGaxO7 (a) x = 0.40, (b) x = 0.50, (c) x = 0.60 and (d) x = \n0.70, registered at T = 5 K, measured after zero field cooling (red open circles), field cooling in \na field of 2 T (black lines) and field cooling in a field of - 2 T (blue lines) . \n \n \n \n \n \n \n \n \n \n \n \n \n \n 22 \n \n \nFig. 7. Temperature dependence of the real (in -phase) component of a. c. susceptibility for \nYBaFe 3.6Ga0.4O7 as a function of temperature measured in zero magnetic field (H dc = 0), using \na frequency of 10 kHz. Inset (a) shows the imaginary (out -of-phase) component of the a. c. \nsusceptibi lity, and inset (b) shows the real (in -phase) component of the a. c. susceptibility \nmeasured using four different frequencies. \n \n \n \n \n \n \n \n \n \n \n \n 23 \n \n \nFig. 8. Temperature dependence of the real (in -phase) component of a. c. susceptibility for \nYBaFe 3.3Ga0.7O7 as a f unction of temperature measured in zero magnetic field (H dc = 0), using \na frequency of 10 kHz. Inset (a) shows the imaginary (out -of-phase) component of the a. c. \nsusceptibility, and inset (b) shows the real (in -phase) component of the a. c. susceptibility \nmeasured using four different frequencies. \n \n \n \n \n \n \n \n \n \n \n \n 24 \n \n \nFig. 9. The (a) real (in -phase) and (b) imaginary (out -of-phase) component of a. c. \nsusceptibility for YBaFe 3.3Ga0.7O7 as a function of temperature. The driving frequency was \nfixed at f = 1 kHz and Hac = 10 Oe. Each curve was obtained under different applied static \nmagnetic field (H dc) ranging from 0 T to 0.2 T. \n \n \n \n \n \n \n \n \n \n \n \n \n 25 \n \n \nFig. 10. The (a) real (in -phase) and (b) imaginary (out -of-phase) component of a. c. \nsusceptibility for YBaFe 3.6Ga0.4O7 as a function of temperature. The driving frequency was \nfixed at f = 1 kHz and H ac = 10 Oe. Each curve was obtained under different applied static \nmagnetic field (H dc) ranging from 0 T to 0.2 T. \n \n \n \n \n " }, { "title": "1210.6700v1.Dirac_half_metal_in_a_triangular_ferrimagnet.pdf", "content": "arXiv:1210.6700v1 [cond-mat.str-el] 24 Oct 2012APS/123-QED\nDirac half-metal in a triangular ferrimagnet\nHiroaki Ishizuka1and Yukitoshi Motome1\n1Department of Applied Physics, University of Tokyo, Hongo, 7-3-1, Bunkyo, Tokyo 113-8656, Japan\n(Dated: September 13, 2021)\nAn idea is proposed for realizing a fully spin-polarized Dir ac semimetal in frustrated itinerant\nmagnets. We show that itinerant electrons on a triangular la ttice exhibit the Dirac cone dispersion\nwithhalf-metallic behavior inthepresence ofathree-subl attice ferrimagnetic order. The Diracnodes\nhavethesamestructureasthoseofgraphene. Byvariational calculation andMonteCarlosimulation,\nwe demonstrate that the ferrimagnetic order with the Dirac n ode spontaneously emerges in a simple\nKondo lattice model with Ising anisotropy. The realization will be beneficial for spintronics as a\ncandidate for spin-current generator.\nPACS numbers: 71.10.Fd, 73.22.Pr, 72.25.-b\nMassless Dirac fermions show substantially different\nnaturefromordinaryelectrons. The peculiarnatureorig-\ninates in the characteristic energy dispersion —the nodal\nstructure with linear dispersion often referred to as the\nDirac cone. While the Dirac fermions were originally\nintroduced in the relativistic quantum theory, recent dis-\ncovery of graphene [1, 2], a single layer sheet of graphite,\nhascarvedoutanewdirectionoftheirstudyincondensed\nmatter systems [3, 4]. In graphene, two Dirac cones ap-\npear in the energy dispersion of πelectrons, which are\nat theKandK′points in the Brillouin zone for the\ntwo-dimensional honeycomb lattice. The Fermi level of\nthis two dimensional conductor comes right at the nodal\npoints, and the low-energy Hamiltonian is well approx-\nimated by the Weyl equation [5]. Various remarkable\nelectronic and transport properties of graphene mainly\nowe to these Dirac cones in the band structure.\nThe extraordinary nature of Dirac fermions in\ngraphene has also attracted a great interest from appli-\ncation to electronics [3]. From the viewpoint of such po-\ntential applications, it is of great interest to control the\ncharacteristic band structure. Furthermore, it is also de-\nsired to control the electronic spin degree of freedom for\nthe application to spintronics [6]. However, there is not\nso much flexibility in graphene, as the Dirac cone is a di-\nrect consequence of the honeycomb lattice geometry and\nthe relativistic spin-orbit interaction is very weak.\nIn this Letter, we propose an alternative solution for\nmanipulating the spin degree of freedom by seeking pos-\nsible emergence of Dirac fermions from itinerant mag-\nnets. We show that itinerant electrons coupled to a\nwell-known ferrimagnet on a triangular lattice give rise\nto the Dirac nodes in their band structure, similar to\nthose of graphene. The resultant masslessDirac fermions\nare spin-polarized, and they are stable in a wide range\nof the spin-charge coupling including typical values in\nsolids. We demonstrate that, by an unbiased Monte\nCarlo (MC) simulation as well as a variational calcu-\nlation, such Dirac half-metal with ferrimagnetic order\nspontaneously emerges in a minimal Kondo-lattice type\nmodel. The results strongly suggest the possibility ofrealizing the exotic electronic state in transition-metal\nand rare-earth compounds, which generally retain much\nhigher controllable degrees of freedom than graphene.\nSuch a new family will not only add a member to the\nknown list of Dirac electrons in solids [7–10], but also\nbring a completely new aspect by the spin polarization.\nIn a half-metal, the electric current is perfectly spin-\npolarized as the low energy excitations only exist for the\nmajority spin [6]. This nature works as a spin-current\ngenerator by filtering out the minority-spin electrons.\nThus, our proposal opens a new frontier for the applica-\ntion of Dirac massless fermions, especially for spintron-\nics [11].\nLet us first discuss a naive, rather trivial approach\nto achieve a Dirac half-metal. We here consider a\nsingle-band ferromagnetic Kondo lattice model (double-\nexchange model) on a honeycomb or kagome lattice [see\nFigs. 1(a) and 1(b)]. The model consists of the nearest-\nneighbor hopping of electrons and the exchange inter-\naction between the electron spin and localized moment,\nwhose Hamiltonian is given by\nH=−t/summationdisplay\n/angbracketlefti,j/angbracketright,σ(c†\niσcjσ+H.c.)−J/summationdisplay\niσi·Si.(1)\nHere,ciσ(c†\niσ) is the annihilation(creation) operatorof an\nitinerant electron with spin σ=↑,↓atith site,σiandSi\nrepresent the itinerant and localized spin, respectively; t\nis the transfer integral and Jis the onsite Kondo cou-\npling. Hereafter we take t= 1 andJ >0.\nIn this model, when Jis sufficiently large compared\nto the bandwidth at J= 0, a ferromagnetic order is\nstabilized by the double-exchange mechanism in a wide\nrange of electron filling n=/summationtext\niσ/angbracketleftc†\niσciσ/angbracketright/2N, whereN\nis the system size [12, 13]. In the ferromagnetic phase,\nthe band is split into two by the large exchange coupling\naccording to the spin component, and each band has ex-\nactly the same form as that for the noninteracting case\nJ= 0. Hence, in principle, the Dirac half-metal arises\nfor the honeycomb and kagome lattices, as the nonin-\nteracting bands on these lattices have the Dirac nodes.\nHowever, these situations are very difficult to realize in2\n(a)\nFIG. 1. (color online). Schematic pictures of (a) a hon-\neycomb ferromagnet, (b) kagome ferromagnet, and (c) three-\nsublattice triangular ferrimagnet. The arrows at each site\nrepresent localized spins.\nsolids as neither such a strong exchange interaction nor\nthe honeycomb and kagome structures is easily realized\nin magnetic compounds.\nAs a more realistic approach, here we propose a sim-\nple, but rather nontrivial route to the half-metallic Dirac\nfermion systems. Let us consider the model in Eq. (1) on\na triangular lattice, and the situation in which a three-\nsublattice collinear ferrimagnetic order with up-up-down\nspin configuration is realized —see Fig. 1(c). By treating\nthe localized moments as classical spins with |Si|= 1,\nthe band structure is easily calculated by the exact diag-\nonalization of the Hamiltonian. The lower three bands\nof the totally six bands are shown in Fig. 2; the two red\nbands are of up spins, and the blue band is of down spin\n(the other upper three bands have the similar form with\nopposite spins).\nThe band structure has a notable feature at the energy\nε=−J; the two up-spin bands touch with each other\nat theKandK′points in the Brillouin zone to form\na Dirac-type point node with linear dispersion, and the\ndown-spinbandhas the bandtop atthe samepoints with\nan ordinary parabolic dispersion. See also the enlarged\nfigure in Fig. 2(b) and the energy dispersion along the\nsymmetric lines in Fig. 2(c).\nInthissituation, whentheelectronfillingisat n= 1/3,\nthe two lower bands are fully occupied while the remain-\ningbands(includingtheupperthree)areunoccupied; theFermi level is located at the nodes where the three bands\nmeet. Asthedown-spinbandhasanenergygap,thehalf-\nmetallic Dirac electrons are obtained by electron doping\nto the unoccupied up-spin band. Although hole doping\nhides the Dirac nature as the down-spin parabolic band\nis doped at the same time, the situation is avoided by in-\ntroducing an additional antiferromagnetic exchange cou-\npling between the neighboring sites, J′/summationtext\n/angbracketlefti,j/angbracketrightσi·Sj[14].\nA finiteJ′>0 shifts the down-spin band to the lower\nenergy and isolates the half-metallic Dirac nodes ener-\ngetically, as demonstrated in Figs. 2(d) and 2(e). Hence,\nthe simple ferrimagnetic order on the triangular lattice\nrealizes the peculiar Dirac half-metallic state near 1/3\nfilling.\nThe Dirac nodes have essentially the same structures\nas those in graphene. Under the ferrimagnetic order, the\nHamiltonian is written as\nH=/summationdisplay\nk\n−Jσz\nAτkτ∗\nk\nτ∗\nk−Jσz\nBτk\nτkτ∗\nk(J+6J′)σz\nC\n.(2)\nHere, the upper tworowscorrespondto the sites with the\nup localized moment ( A,Bsublattices) and the bottom\nrow is for the down one ( Csublattice) in the three-site\nunit cell. In Eq. (2), σzis thezcomponent of the Pauli\nmatrix for itinerant electrons, kis the wave vector, and\nτkis the Fourier transform of the hopping term given by\nτk=−t[eikx+ei/parenleftBig\n−kx\n2+√\n3\n2ky/parenrightBig\n+ei/parenleftBig\n−kx\n2−√\n3\n2ky/parenrightBig\n]. By using\nthek·pperturbation around the KandK′points in the\nBrillouin zone [5] and by expanding the result up to the\nfirst order in terms of tκx/Jandtκy/J(κis the relative\nwave vector measured from KandK′points), we end up\nwith the low-energy Hamiltonian which is factorized into\ntwo parts. One is a 2 ×2 Hamiltonian for the up-spin\nhoneycomb subnetwork of the AandBsublattices, and\nthe other is a localized state at the down-spin sites in the\nCsublattice. The former is given by\nHDirac\nk±=/parenleftbigg\n−J3\n2it(κx±iκy)\n−3\n2it(κx∓iκy)−J/parenrightbigg\n,(3)\nwhere the sign ±corresponds to the KandK′points.\nThis has an equivalent form to that of graphene.\nIt is worthy to note that the Dirac nodes are formed\nimmediately by switching on J. However, when Jis\nsmall, the low-energy physics at n= 1/3 is not char-\nacterized solely by the massless Dirac fermions because\nthere is a band overlap at the energy of the Dirac\nnodes. The band overlap comes from the second lower\nband for up spin, which has an energy minimum at\nk= (2π/3,0) points and its threefold symmetric points\nfor smallJ; the minimum energy is given by ε(2π\n3,0)=\nt/2−/radicalig\n(J+3J′−t/2)2+2t2. In order for the Dirac\nnodes to be isolated at the Fermi level, this energy\nshould be higher than that at the KandK′points,3\nKM Γ Γ01\n-1\n-2\n-3\n-4\n-5\n-6\n-7(c)\nε(a)\n2\n0\n-2\n-4\n-6\n0\n2π\n32π\n3-\n04π\n3√3-ε\nkx\nkyK\nMΓ\n-8\nK'\n4π\n3√3(b)\nε\nkykx-1.0\n-1.5\n-2.0\n-2.5\n-3.0\n00(d)\nε\nkykx-1.0\n-1.5\n-2.0\n-2.5\n-3.0\n00\n(e)\nKM Γ Γ01\n-1\n-2\n-3\n-4\n-5\n-6\n-7ε2π\n34π\n3√32π\n34π\n3√3\nFIG. 2. (color online). Band structures of the model in Eq. (1 ) under the three-sublattice ferrimagnetic order at J= 2. (a)\nThe overall band structure of the three lower-energy bands a tJ′= 0, (b) the enlarged view near the Fermi level ε=−Jat\nn= 1/3 in the first quadrant, and (c) the cut along the symmetric lin es. (d) and (e) show the results at J′= 0.05. The arrows\nindicate the spins for each band. In (a), the gray hexagon on t he basal plane shows the first Brillouin zone for the magnetic\nsupercell. The dashed line in (e) indicates the Fermi level i n the MC simulation shown in Fig. 4.\nεK=−(J+3J′). Hence, the Dirac nodes are energeti-\ncally isolated and play a decisive role when the condition\n(J+3J′)/t>1 is satisfied. This condition is important\nbecausethenecessary JandJ′aremuchsmallerthanthe\nnoninteracting bandwidth 9 t, and it is indeed satisfied in\nwide range of materials.\nSo far, we assumed the presence of three-sublattice fer-\nrimagnetic order. In the following, we show that such\norder is indeed stable in the Kondo-lattice type model\nas Eq. (1). We here simplify the model by assuming the\nlocalized moments are the Ising spins taking the values\nSi=±1.\nFirst, we investigate the ground state phase diagram\nnearn= 1/3 by a variational calculation. We com-\npare the ground state energy of the two-sublattice stripe\nphase and three-sublattice ferrimagnetic phase appeared\nin the previous study [15], in addition to the ferromag-\nnetic phase. The results at J= 2 are shown in Fig. 3 for\nJ′= 0 and 0.05. AtJ′= 0, the ground state in the plot-\nted rangeis dominated by the ferrimagneticphaseas well\nas the stripe phase. The different phases are separated\nbyphaseseparation. As showninFig. 3(b), the introduc-\ntion of small J′largely stabilizes the ferrimagnetic phase\nnearn= 1/3 as well as the stripe phase. This is because\nthe itinerant electron spins are polarized parallel to the\nlocalized spins in the ground state, leading to an energy\ngain (loss) by the antiferromagnetic J′for the two states\n(the ferromagnetic state).\nWe next examine the stability of the ferrimagnetic or-\nderatfinite temperaturesbyanunbiasedMC simulation.\nFor the simulation, a standard algorithm for fermionnJ\n012345678\n0.260.280.30 0.38 0.320.340.36\nnJ\n2-sub stripe\n0.260.280.30 0.38 0.320.340.36012345678(a)\nFIG. 3. (color online). Ground state phase diagram obtained\nby variational calculation at (a) J′= 0 and (b) J′= 0.05.\nThe schematic picture of magnetic structure in each phase\nis shown. The white region indicates the electronic phase\nseparation (PS)andthedottedvertical lines indicate n= 1/3.4\nMxy[1/√2 ]\n|Mz|[x√3]\nN =12x12\nN =12x18\nN =18x18\n0.00.20.40.60.81.01.2\n(a)Tc\nTN =12x12\nN =12x18\nN =18x18\n0.00.20.40.60.8\n0.00 0.05 0.10 0.15 0.20 0.25(c)\nψχz[x5] N =12x12\nN =12x18\nN =18x18\n020406080100120140160180(b)\nχxyTKT\nFIG. 4. (color online). MC results for (a) the pseudo mo-\nmentsMxyand|Mz|, (b) corresponding susceptibilities χxy\nandχz, and (c) azimuth parameter ψ. The data are calcu-\nlated atn= 0.34.\nsystems coupled to classical fields is used [16]. In this\nmethod, the trace overthe fermions in the partitionfunc-\ntion is calculated by the exact diagonalization, while the\ntrace over classical spin configurations is computed by\na classical MC method using the Metropolis dynamics.\nThe phase transition to ferrimagnetic phase is detected\nby using two parameter [17]. One is the pseudo-moment\ndefined by\n˜Sm=\n2√\n6−1√\n6−1√\n6\n01√\n2−1√\n21√\n31√\n31√\n3\n\nSi\nSj\nSk\n,(4)\nwheremis the index for the three-site unit cells, and\n(i,j,k) denote the three sites in the mth unit cell be-\nlonging to the sublattices ( A,B,C), respectively. We\nmeasure the summation M= (3/N)/summationtext\nm˜Smand the\nsusceptibility. The other is the azimuth parameter ψ\ndefined by ψ= (˜Mxy)3cos6φM, whereφMis the az-\nimuth angle of Min thexyplane and ˜Mxy= 3M2\nxy/8\n(M2\nxy=M2\nx+M2\ny). The ferrimagnetic ordering is sig-\nnaled byMxy→2/radicalbig\n2/3,|Mz| →1/√\n3, andψ→1 at\nlow temperature T→0, respectively [15, 18, 19].\nFigure 4 shows the MC results at J= 2 andJ′= 0.05\nin the slightly electron doped region to n= 1/3 [see alsoFig.2(e)]. Theresultsindicatetwosuccessivephasetran-\nsitions atTKT= 0.192(15) and at Tc= 0.108(9). The\ntransition temperatures are estimated by extrapolating\nthe peak of susceptibilities χxyandχzasN→ ∞. The\ntransition at TKTis considered as a Kosterlitz-Thouless\ntype with the growth of quasi-long-range order [15]. On\nthe other hand, the phase transition at Tcis a three-\nsublatticeferrimagneticordering. TheMC resultand the\nabove analysis for the ground state consistently indicate\nthat the three-sublattice ferrimagnetic order is stabilized\nin the vicinity of n= 1/3 in the wide range of parame-\nters forJandJ′, spontaneously giving rise to the Dirac\nhalf-metal.\nAs such ferrimagnetic order was indeed observed in\nseveralinsulatingmagnets[20,21], ourresultsinthemin-\nimal model will stimulatethe hunt forDirachalf-metal in\ntransition-metal and rare-earth compounds. The present\nresults will be qualitatively robust even when extend-\ning the model to more realistic situation. For instance,\ntheferrimagneticstateremainsstablewhenincluding the\ntransverse components of localized spins, at least, in the\npresence of the Ising anisotropy. Multi-band effect may\nbe avoided under a particular crystal field; for instance,\nthed-electrona1gorbital isolated by a strong trigonal\nfield is a good candidate for the realization. Interlayer\ncoupling, however, may open a gap at the Dirac nodes.\nNevertheless, a straightforward stacking of layers or suf-\nficiently isolated layers in a controlled thin film will be\npromising to preserve the massless nature.\nThe authors thank Y. Matsushita, A. Shitade, and\nY. Yamaji for helpful comments. H.I. is supported\nby Grant-in-Aid for JSPS Fellows. This research\nwas supported by KAKENHI (No.19052008, 21340090,\n22540372, and 24340076), Global COE Program “the\nPhysical Sciences Frontier”, the Strategic Programs for\nInnovative Research (SPIRE), MEXT, and the Compu-\ntational Materials Science Initiative (CMSI), Japan.\n[1] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,\nY. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A.\nFirsov, Science 306, 666 (2004).\n[2] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,\nI. V. Katsunelson, I. V. Grigorieva, S. V. Dubonos, and\nA. A. Firsov, Nature 438, 197 (2005).\n[3] A. K. Geim and K. S. Novoselov, Nature Mater. 6, 183\n(2006).\n[4] For a recent review, see A. H. Castro Neto, N. M. R.\nPeres, K. S. Novoselov, and A. K. Geim, Rev.Mod. Phys.\n81, 109 (2009).\n[5] G. W. Semenoff, Phys. Rev. 53, 2449 (1984).\n[6] S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J.\nM. Daughton, S. von Molnar, M. L. Roukes, A. Y.\nChtchelakanova, amd D. M. Treger, Science 294, 1488\n(2001).\n[7] M. H. Cohen and E. I. Blount, Phil. Mag. 5, 115 (1960).5\n[8] T. Konoike, K. Uchida, and T. Osada, J. Phys. Soc. Jpn.\n81, 043601 (2012).\n[9] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045\n(2010).\n[10] S. Ishibashi, K. Terakura, and H. Hosono, J. Phys. Soc.\nJpn.77, 053709 (2008).\n[11] D. Pesin and A. H. MacDonald, Nature Mater. 11, 409\n(2012).\n[12] C. Zener, Phys. Rev. 82, 403 (1951).\n[13] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675\n(1955).\n[14] An off-site Kondo coupling may exist generally in the\nKondo lattice systems, although the magnitude is much\nsmaller than the onsite one and the sign depends on\nthe orbital nature of itinerant and localized electrons.\nThe antiferromagnetic superexchange coupling between\nneighboring localized spins, given by JAF/summationtext\n/angbracketlefti,j/angbracketrightSi·Sj,\nmay also exist, but it neither modifies the band structure\nnor harms the stability of the ferrimagnetic state.[15] H. Ishizuka and Y. Motome, Phys. Rev. Lett. 108,\n257205 (2012).\n[16] S. Yunoki,J. Hu, A.L.Malvezzi, A.Moreo, N.Furukawa,\nand E. Dagotto, Phys. Rev. Lett. 80, 845 (1998).\n[17] In principle, the ferrimagnetic ordering can be detect ed\nby the spin structure factor. In the finite-size MC cal-\nculations for the current model, however, it is useful to\nemploy the pseudo-moment and its azimuth parameter\nfor distinguishing it from a three-sublattice partial diso r-\nder and Kosterlitz-Thouless type quasi long-range order.\nSee also Ref. [15, 18, 19].\n[18] H. Takayama, K. Matsumoto, H. Kawahara, and K.\nWada, J. Phys. Soc. Jpn. 52, 2888 (1983).\n[19] S. Fujiki, K. Shutoh, S. Inawashiro, Y. Abe, and S. Kat-\nsura, J. Phys. Soc. Jpn. 55, 3326 (1986).\n[20] M. Tanaka, H. Iwasaki, K. Siratori, and I. Shindo, J.\nPhys. Soc. Jpn. 58, 1433 (1989).\n[21] J. Iida, M. Tanaka, Y. Nakagawa, S. Funahash, N.\nKimizuka, and S. Takekawa, J. Phys. Soc. Jpn. 62, 1723\n(1993)." }, { "title": "1010.4872v1.High_spin_polarization_in_epitaxial_films_of_ferrimagnetic_Mn3Ga.pdf", "content": "High spin polarization in epitaxial films of ferrim agnetic Mn 3Ga \n \nH. Kurt,* K. Rode. M. Venkatesan, P. S. Stamenov an d J. M. D. Coey \n \nSchool of Physics and CRANN, Trinity College, Dubli n 2, Ireland \n \n* kurth@tcd.ie \n \nPACS number(s): 75.76.+j, 75.70.-i, 75.60.Ej, 75.50 .-y \n \nAbstract \n \nFerrimagnetic Mn 3Ga exhibits a unique combination of low saturation magnetization \n(Ms = 0.11 MA m -1) and high perpendicular anisotropy with a uniaxial anisotropy \nconstant of Ku = 0.89 MJ m -3. Epitaxial c-axis films exhibit spin polarization as high \nas 58%, measured using point contact Andreev reflec tion. These epitaxial films will \nbe able to support thermally stable sub-10 nm bits for spin transfer torque memories. \n \n \n \n \n \n \n \n There is a revival of interest in Heusler alloys, m otivated by the extraordinary \nvariety of electronic ground states that can be ach ieved by varying the number of \nvalence electrons 1, 2, composition 3-5 and atomic order in these materials. The Heusler \nfamily include semiconductors, metals and half meta ls 6, as well as ferromagnets, \nantiferromagnets and superconductors, and even perh aps compensated ferrimagnetic \nhalf metals 7 and topological insulators 8, 9. Thin films of half-metallic Heuslers with a \nhigh Curie temperature, such as Co 2MnSi, have been successfully used in magnetic \ntunnel junctions 10-13 and spin valves 14 . The magnetization of these films lies in-plane, \nbut tetragonally-distorted Heuslers could offer hig h perpendicular anisotropy, \nnecessary for thermally-stable sub-10 nm tunnel jun ctions and spin valves. Here we \npresent epitaxially grown tetragonal Mn 3Ga films, that exhibit a spin polarization of \nup to 58%, together with uniaxial anisotropy ( K1 = 0.89 MJ m -3) and low \nmagnetization ( Ms = 0.11 MA m -1), a combination of properties that may be ideal fo r \ntiny perpendicular spin-torque switchable elements that will allow for scalable \nmagnetic memory and logic. \n \nThe general composition of the L21 cubic Heusler alloys is X 2YZ. In the \nperfectly-ordered state, illustrated in Fig. 1a, th e Y and Z atoms occupy two \ninterpenetrating face-centred cubic lattices, where each is octahedrally coordinated by \nthe other, and the X atoms form a simple cubic latt ice, where they are tetrahedrally \ncoordinated by both Y and Z atoms, at the corners o f a cube. The X–Y , X–X, Y–Y \nbond lengths are √3a0/4, a0/2 and a0/√2, respectively, where a0 is the cubic lattice \nparameter. The material we discuss here, Mn 3Ga, forms two stable crystal structures. \nThe high temperature hexagonal D019 phase is a triangular antiferromagnet, easily \nobtained by arc melting 15, 16 . The tetragonal D022 phase is a ferrimagnet, usually obtained by annealing the hexagonal material at 350 -400 °C for 1-2 weeks 15, 17, 18 . The \nspin polarization at the Fermi level has been calcu lated to be 88 % for the tetragonal \nphase, which has been suggested as a potential mate rial for spin torque applications 19 . \nTo this end, thin films with appropriate magnetic p roperties were needed. \n \nThe D022 structure is a highly-distorted tetragonal variant of the L21 Heusler \nunit cell, which has been stretched by 27% along c axis. The unit cell is outlined on \nFig. 1a by the red line; lattice parameters of the unit cell (space group I4/ mmm ), \nwhich contains two Mn 3Ga formula units, are a = 394 pm and c = 710 pm. As a result \nof the tetragonal structure, both 4 d tetrahedral X-sites and 2b octahedral Y-sites are \nsubject to strong uniaxial ligand fields, which lea d to uniaxial anisotropy at both Mn \npositions. Imperfect atomic order results in some M n population of the 2 a Z sites, \nwhich are similarly distorted. \n \nGenerally, the Mn moment and Mn-Mn exchange in meta llic alloys are very \nsensitive to the inter-atomic distances. Widely-spa ced Mn atoms with a bond length ≥ \n290 pm tend to have a large moment, of up to 4 µB, and couple ferromagnetically 20 . \nNearest-neighbour Mn atoms have much smaller moment s, and couple \nantiferromagnetically when the bond length is the r ange 250 – 280 pm. As a result, the \nmagnetic order is often complex. αMn, for example has a large cubic cell ( a = 890 \npm) with four different manganese sites having mome nts ranging from 0.5 to 2.6 µB, \nand a complex noncollinear antiferromagnetic struct ure with TN = 95 K. The \nmagnetism of tetragonal Mn 3Ga is simpler. The Y sublattice is ferromagneticall y \ncoupled, as expected from the long Y-Y bonds (391, 450 pm), but there is a strong \nantiferromagnetic X-Y intersublattice interaction ( bond length 264 pm) which leads to ferrimagnetic order. This overcomes the X-X interac tions, which are \nantiferromagnetic in-plane (bond length 278 pm) and ferromagnetic along the c-axis \n(bond length 355 pm). The antiferromagnetic X-X int eractions are frustrated by the \nstrong antiferromagnetic X-Y coupling, and the X su blattice remains ferromagnetic, \nwith its moment aligned antiparallel to that of the Y sublattice. Different occupancies \nand magnetic moments in X and Y sites leads to a co mpensated ferrimagnetic \nstructure with alternating spins on atomic planes a long c axis (Fig. 1b). Imperfect \natomic ordering of Mn and Ga may, however, produce local deviations from collinear \nferrimagnetism. The Curie temperature of Mn 3Ga is high. It would be much greater \nthan 770 K were it not for the transition to the an tiferromagnetically ordered \nhexagonal D019 structure at this temperature 15, 19 . \n \nFig. 1 (a) The cubic unit cell of the L21 Heusler alloys. The tetragonal unit cell \nof the D022 structure is indicated by the red line. (b) In Mn 3Ga, it is stretched along \nthe c-axis. The X and Y positions are occupied by Mn, an d the Z position is occupied \nby Ga. The magnetic couplings in X and Y sublattice s are antiferromagnetic and \nferromagnetic respectively, but the stronger antife rromagnetic intersublattice coupling \nbetween X and Y sites frustrates the X site moments leading to a ferrimagnetic structure. \n \nWe have grown oriented films of Mn 3Ga on MgO (001) substrates with various buffer \nlayers by dc magnetron sputtering from a Mn 3Ga target in a chamber with 2.0 × 10 -8 \nTorr base pressure. The Mn 3Ga growth rate was approximately 1nm/min. The \nsubstrate temperature was varied from 250-375 °C. Buffer layers were 10-30 nm of \noriented Pt (001), Cr (001) or MgO (001). Pt seed l ayers were prepared by pulsed \nlaser deposition at 500 °C substrate temperature under 40 µbar oxygen partia l \npressure. The films grow with (001) orientation ( c-axis normal to plane) on all three. \nMost of our stacks were made by dc-magnetron sputte ring; Mn 3Ga films grown by \npulsed laser deposition turned out to have the hexa gonal D019 structure. X-ray \ndiffraction was carried out using a Bruker D8 Disco very diffractometer with a \nmonochromated Cu K α1 source. Magnetization measurements were made usin g a \nQuantum Design MPMS SQUID and PPMS vibrating sample magnetometers. \n \nStructural, magnetic and spin polarization data for representative films are \nsummarized in Table 1, and some X-ray diffraction d ata are shown in Fig. 2a-c. \nSample 345 was a perpendicular spin valve structure with two Mn 3Ga layers, \nseparated by a 10 nm Cr spacer. The relatively high magnetization and low spin \npolarization of the Mn 3Ga layer grown on Cr indicate that it has a high de gree of \ndisorder. \n \nFig. 2. a) X-ray diffraction patterns of Mn 3Ga layers grown on Pt, Cr and MgO seed \nlayers. b) Substrate temperature ( Ts) dependence of Mn 3Ga layers grown on Pt (001); \nthe inset shows the order parameter S vs. Ts, at Ts = 375 °C the sample is not single \nphase. c) φ scans showing the four-fold symmetry of the films a nd the MgO (001) \nsubstrate. d) Atomic force microscope images of fil ms grown on MgO, e) Pt, f) Cr, all \ngrown at Ts = 350 °C. \n \n \n \nTable 1. Structural, magnetic and spin polarization data on Mn 3Ga films. Films grown \non well oriented Pt and MgO exhibit lower magnetiza tion accompanied with a small \nin-plane canted moment and higher spin polarization , whereas the films grown on Cr \nhave higher moment and reduced polarization. Sample 345a and b represent a bilayer \nMn 3Ga structure; MgO/Pt/Mn 3Ga(60)/Cr(10)/Mn 3Ga(55). \n \nSample Buffer Ts t c M // M ⊥ m µ0Hc P (300K) (300K) (300K) (2.2K) \n# °C nm pm MA m -1 MA m -1 µB fu -1 T % \n343 MgO/Pt 250 90 707 0.022 0.11 0.65 1.80 45 \n346 MgO/Pt 250 90 712 0.017 0.10 0.59 1.90 55 \n388 MgO 350 60 707 0.015 0.14 0.81 1.80 n.a. \n408 MgO/Cr 350 40 696 0 0.14 0.81 1.68 n.a. \n345a MgO/Pt 250 60 708 n.a. 0.15 0.89 1.50 58 \n345b Cr 250 55 700 n.a. 0.2 2 1.3 0 1.02 40 \n* 1 MA m -1 corresponds to 5.93 µB fu -1 \n \n \n \nThe composition of one of the films grown on MgO wa s determined by ICP \nmass spectroscopy; the Mn:Ga atomic ratio was 3.0:1 .0. The films were metallic, and \nthe resistivity of those grown on Pt was estimated as 1.6 ×10 -6 Ω m after correcting \nfor the resistance of the underlying Pt. Resistivi ty of a film grown directly on MgO \ncould not be measured because of the discontinuous island growth mode in this case, \nillustrated in Fig. 2d. \n \nAn order parameter S is defined by the square root of the intensity rat io of \n(101) to (204) reflections, divided by the theoreti cal ratio. The variation of S with \nsubstrate temperature is shown in the inset of Fig. 2b. The highest value of 0.8 is \nmeasured for the samples grown on Pt (001) at 350 °C. Platinum has the smallest \nlattice mismatch with tetragonal Mn 3Ga (0.4%) and almost strain-free epitaxial \ngrowth takes place easily. The surfaces of Mn 3Ga films exhibit variations depending on the substrate type and growth temperature as sho wn in Fig. 2d-f. The pinhole-free \nareas of films grown on Pt seed layers at 350 °C have an rms surface roughness of 0.8 \nnm, whereas the samples grown on thin Cr seed layer s were pinhole free, but much \nrougher (4-5 nm rms). In addition, there are large, randomly distributed islands ~ 50 \nnm high on the surface (Fig. 2f). Moreover, the Mn 3Ga films grown on Cr have a \nshorter c-axis, which is probably caused by the large lattic e mismatch with Cr (4.2 \n%). The D022 unit cell has to expand in the ab - plane for epitaxial growth. The films \non Pt and MgO grow in cube on cube mode (Fig. 2c), whereas the films on Cr are \nrotated in-plane by 45 degrees to facilitate epitax ial growth. The reason for the island \ngrowth mode on MgO is thought to be the large latti ce mismatch (7.7%), as islands \ncan relax the lattice strain. The top surface of th ese films is smooth, but continuous \nfilm growth was not possible for thicknesses up to 60-70 nm. The pinholes seen on \nthe surface of Mn 3Ga grown on Pt simply reflect pinholes formed on th e surface of \nthe thin Pt (001) seed layer, which arise because o f the lattice mismatch between Pt \nand MgO. Pinhole free surfaces can be obtained on t hick Pt (001) seedlayers (> 200 \nnm) grown on MgO (001) substrates, which show feat ureless surfaces 21 . \n \nThe relatively high surface roughness of epitaxial Mn 3Ga may limit its use to \ngiant magnetoresistance (GMR) spin valve devices. T his may not be a disadvantage. \nAs the dimensions of the memory bits decreases belo w 50 nm the large impedance of \ntunnel junctions can limit their use in high-speed circuits, whereas the impedance of \nGMR devices based on Mn 3Ga could be engineered to 50 Ω in these dimensions, \nwhich is compatible with high-speed operation. \n \nA typical room-temperature hysteresis loop is shown in Fig. 3a. Data are corrected for substrate diamagnetism. The substrate also exhibits weak Curie-law \nparamagnetism at low temperatures, which is attribu ted to 50 ppm of Fe 2+ impurities \npresent in the MgO. The broad hysteresis loops show a spontaneous magnetization of \n130 ± 20 kA m -1 and coercivity of up to 1.9 T at 300 K. Epitaxial Mn 3Ga films grown \non MgO and Pt also exhibit a small canted magnetic moment which appears in the \nmagnetization measurements with field parallel to t he plane. There is considerable \nvariation in coercivity, but similar values of magn etization are found in all samples, \ncorresponding to a moment of 0.59 – 0.89 µB fu -1. Values reported in bulk samples \ndepend on the stoichiometry of the D022 compound, increasing from about 1.0 µB fu -1 \nfor a material with a 3:1 Mn:Ga ratio to 1.4 µB fu -1 for a 2:1 ratio 16-19 . The \nmagnetization in thin films is similar, and it has been shown to depend on the degree \nof atomic ordering 22, 23 . The magnetization is consistent with an essential ly \nantiparallel arrangement of moments on the 4 d and 2 b sites, determined by neutron \ndiffraction for a sample of composition Mn 2.85 Ga 1.15 to be 1.6 ± 0.2 and -2.8 ± 0.3 µB, \nrespectively 15 . A higher moment in the 2 b site is consistent with the calculations of \nKübler et al .24 . Electronic structure calculations predict larger moments on both sites, \nand indicate that the compound is nearly half-metal lic with a spin polarization of 88 \n% at the Fermi level 19 . \n \nThe strength of the uniaxial anisotropy can best be determined from the \nperpendicular and parallel magnetization curves. Th e anisotropy field determined \nfrom the extrapolation in Fig. 3a is 16 T, which co rresponds to a uniaxial anisotropy \nconstant K1 = 0.89 MJ m -3. A somewhat larger value was reported by Wu et al . in \nMn 2.5 Ga, which is mainly due to the higher magnetization on account of non-\nstoichiometric composition 23 . The inverse correlation of coercivity and magneti zation indicated in Table 1 reflects the variation of the anisotropy field Ha = 2 K1 /Ms. The \nsmallest volume V for which the stability condition KuV/kBT ≥ 60 is satisfied at room \ntemperature is 280 nm 3. The corresponding nanopillar with height equal to diameter \nhas dimensions 7 nm. \n \nIn uniformly magnetized, homogeneous uniaxial magne ts, the anisotropy field \nsets an upper limit on the coercivity as pointed ou t by Stoner and Wohlfarth 25 . In \nreality, the coercivity is always much lower, a res ult known as Brown’s paradox 26 . \nThe explanation is that material is never perfectly homogeneous; reversal begins in a \nlocalized nucleation volume of δB3, where δB = π √(A/Ku) is the Bloch wall width, and \nA is the micromagnetic exchange constant. Assuming A ≈ 10 pJ m -1, we find δB = 10 \nnm. The expected activation volume is of order 100 0 nm 3. We anticipate that \nnanostructured Mn 3Ga elements will exhibit stable coherent reversal a t dimensions \nwhich are less than 10 nm. It therefore promises sc aling to these sizes. \n \nMultilayer spin valve stacks were realized by growi ng the second layer of Mn 3Ga \non a Cr spacer. The magnetization measurements reve al a two step switching curve \nfor the trilayer structure as shown in Fig. 3b. The magnetization reversal of the film \ngrown on Cr is sharper with a coercivity of 1T, whe reas the film grown on Pt switches \nat 1.5 T. The highest measured spin polarization of the Mn 3Ga films is 58 %, which \ncompares with the calculated value of 88% 19 . The polarization decreases with order, \nstrain as well as Mn deficiencies, which increases magnetization. \n \n \nFig. 3. (a) Magnetization curves for a 90 nm thick film of tetragonal Mn 3Ga measured at room \ntemperature with the field applied perpendicular an d parallel to the plane of the film. Anisotropy fie ld \nat 300 K was determined by extrapolating in plane d ata through origin and is 16 T. (b) Magnetization \nof Mn 3Ga/Cr/Mn 3Ga bilayer structure. (c) Point-contact Andreev ref lection spectrum of a Mn 3Ga film \ngrown on Pt. \n \n \nThe tetragonal D022 -Mn 3-xGa (x = 0 - 1) offers a wide variety of magnetic \nproperties that can be engineered to suit a specifi c application. In the case of \nstoichiometric Mn 3Ga, presented here, the magnetization is reduced du e a higher \ndegree of compensated spins and high anisotropy wou ld allow thermally-stable sub-\n10 nm spin torque devices. For magnetic random acce ss memories, a low \nmagnetization has the advantage that it requires a lower current to switch by STT. A \nparticular advantage is that the Mn 3Ga nanopillars would be switchable by STT but \nimmune to magnetic field of even NdFeB permanent ma gnets due to the high \ncoercivity of the films, and therefore would not re quire any magnetic shielding. \nIn conclusion, tetragonal Mn 3Ga thin films look promising for nanoscale spin-\ntransfer torque memory and logic applications. Ther e is sufficient anisotropy to ensure \nthermal stability to sub-10 nm dimensions, and the small magnetization is \nadvantageous for spin torque switching. There is a degree of flexibility in the D022 \nstructure, in terms of composition and degree of at omic order, which should enable \nthe magnetic properties to be optimized for a speci fic magnetic application. The immediate challenge now is to observe spin torque s witching a Mn 3Ga nanopillar. \n \n \n \n \nAcknowledgements \nThis work was supported by SFI as part of the MANSE project 2005/IN/1850, and \nwas conducted under the framework of the INSPIRE pr ogramme, funded by the Irish \nGovernment's Programme for Research in Third Level Institutions, Cycle 4, National \nDevelopment Plan 2007-2013. \n \n \n \n \n \n1 I. Galanakis, P . Mavropoulos, and P . H. Dederichs, Journal of Physics D: Applied Physics 39 , \n765 (2006). \n2 I. Galanakis, P . H. Dederichs, and N. Papanikolaou, Physical Review B 66 , 174429 (2002). \n3 P . J. Webster and K. R. A. Ziebeck, in Landolt-Börnstein, New Series, Group III , edited by W. H. \nR. J. (Springer, Berlin, 1988), Vol. 19c, p. 77 \n4 H. C. Kandpal and et al., Journal of Physics D: App lied Physics 40 , 1507 (2007). \n5 B. Hillebrands and C. Felser, Journal of Physics D: Applied Physics 39 (2006). \n6 R. A. de Groot, F. M. Mueller, P . G. v. Engen, and K. H. J. Buschow, Physical Review Letters 50 , \n2024 (1983). \n7 S. Wurmehl, H. C. Kandpal, G. H. Fecher, and C. Fel ser, Journal of Physics Condensed Matter \n18 , 6171 (2006). \n8 S. Chadov, X. Qi, J. Kübler, G. H. Fecher, C. Felse r, and S. C. Zhang, Nat Mater 9, 541 (2010). \n9 H. Lin, L. A. Wray, Y . Xia, S. Xu, S. Jia, R. J. Ca va, A. Bansil, and M. Z. Hasan, Nat Mater 9, 546 \n(2010). \n10 W. Wang, H. Sukegawa, R. Shan, S. Mitani, and K. In omata, Applied Physics Letters 95 , 182502 \n(2009). \n11 S. Tsunegi, Y . Sakuraba, M. Oogane, K. Takanashi, a nd Y . Ando, Applied Physics Letters 93 , \n112506 (2008). \n12 N. Tezuka, N. Ikeda, S. Sugimoto, and K. Inomata, A pplied Physics Letters 89 , 252508 (2006). \n13 Y . Sakuraba, M. Hattori, M. Oogane, Y . Ando, H. Kat o, A. Sakuma, T. Miyazaki, and H. Kubota, \nApplied Physics Letters 88 , 192508 (2006). \n14 T. Iwase, Y . Sakuraba, S. Bosu, K. Saito, S. Mitani , and K. Takanashi, Applied Physics Express 2, \n063003 (2009). \n15 E. Krén and G. Kádár, Solid State Communications 8, 1653 (1970). \n16 H. Niida, T. Hori, and Y . Nakagawa, Journal of the Physical Society of Japan 52 , 1512 (1983). \n17 H. Niida, T. Hori, H. Onodera, Y . Yamaguchi, and Y . Nakagawa, Journal of Applied Physics 79 , \n5946 (1996). \n18 B. Balke, G. H. Fecher, J. Winterlik, and C. Felser , Applied Physics Letters 90 , 152504 (2007). 19 J. Winterlik, B. Balke, G. H. Fecher, C. Felser, M. C. M. Alves, F. Bernardi, and J. Morais, \nPhysical Review B 77 , 054406 (2008). \n20 N. Yamada, Journal of the Physical Society of Japan 59 , 273 (1990). \n21 G. Cui, V. Buskirk, J. Zhang, C. P . Beetz, J. Stein beck, Z. L. Wang, and J. Bentley, in Materials \nResearch Society Symposium Proceedings 310 , 1993), p. 345. \n22 F. Wu, E. P . Sajitha, S. Mizukami, D. Watanabe, T. Miyazaki, H. Naganuma, M. Oogane, and Y . \nAndo, Applied Physics Letters 96 , 042505 (2010). \n23 F. Wu, S. Mizukami, D. Watanabe, H. Naganuma, M. Oo gane, Y . Ando, and T. Miyazaki, Applied \nPhysics Letters 94 , 122503 (2009). \n24 J. Kübler, A. R. William, and C. B. Sommers, Physic al Review B 28 , 1745 (1983). \n25 E. C. Stoner and E. P . Wohlfarth, Philosophical Tra nsactions of the Royal Society of London. \nSeries A, Mathematical and Physical Sciences 240 , 599 (1948). \n26 W. F. Brown, Reviews of Modern Physics 17 , 15 (1945). \n \n " }, { "title": "1209.0004v1.Magnetic_symmetry_of_the_plain_domain_walls_in_the_plates_of_cubic_ferro__and_ferrimagnets.pdf", "content": "1 \n \n MAGNETIC SYMMETRY OF THE PLAIN DOMAIN WALLS IN THE PLATES OF \nCUBIC FERRO- AND FERRIMAGNETS \nB. M. Tanygin 1, O. V. Tychko 2 \nKyiv Taras Shevchenko National University, Radiophy sics Faculty, Volodumiurska 64, \nKyiv, MSP 01601, Ukraine \n1E-mail: b.m.tanygin@gmail.com, 2E-mail: a.tychko@gmail.com \nAbstract. Magnetic symmetry of possible plane domain walls i n arbitrary oriented plates of \nthe crystal of hexoctahedral crystallographic class is considered. The symmetry classification \nis applied for ferro- and ferrimagnets. \nPACS: 61.50 Ah, 75.60 Ch \n1. Introduction \nFor sequential examination of static and dynamic pr operties of domain walls (DWs) \nin magnetically ordered media it is necessary to ta ke into account their magnetic symmetry \n[1,2]. The complete symmetry classification of plan e 180°-DWs in magnetically ordered \ncrystals [1], similar classification of these DWs w ith Bloch lines in ferromagnets and ferrites \n[2] and magnetic symmetry classification of plane n on 180 0-DWs (all possible DW types \nincluding 0 0-DWs [3]) in ferro- and ferrimagnets [4] were carri ed out earlier. These DW \nsymmetry classifications allows arbitrary crystallo graphic point symmetry group of the \ncrystal. The influence of the spatially restricted sample surfaces on the DW magnetic \nsymmetry wasn’t considered in works [1-4]. The real magnetic sample restricts a spatial (3D) \nmagnetization distribution. Therefore, it modifies the DW symmetry in general case. This \npaper presents the investigation of the influence o f the restricted sample surfaces on the \nsymmetry of the all possible (0 0-, 60 0-, 70.5 0-, 90 0-, 109.5 0-, 120 0- and 180 0-DW [4,5]) plane \n(i.e. DW with 0r>> δ, where 0r is the curvature radius of the DW [1]) DWs in an a rbitrary \noriented plate of the cubic (crystallographic point symmetry group m3m) ferro- and \nferrimagnets. \n 2 \n \n 2. Domain wall symmetry in the restricted sample \nThe DW symmetry can be described by the magnetic sy mmetry classes (MSCs) kG \nwhere k is a MSC number [1]. The MSC kG of DW is the magnetic symmetry group \nincluding all symmetry transformations (all transla tions are considered as unit operations) that \ndo not change the spatial distribution of magnetic moments in the crystal with DW. The \nabove-mentioned group is a subgroup of the magnetic (Shubnikov’s) symmetry group ∞\nPG of \nthe crystal paramagnetic phase [6]. Total number of MSCs of arbitrary type DWs (i.e. DWs \nwith arbitrary α2 angle (0 0≤α2≤180 0) between the unit time-odd axial vectors 1m and 2m \ndirected along magnetization vectors 1M and 2M in neighboring domains) in ferro- and \nferrimagnets is equal to 64. General enumeration of MSCs contains 42 MSCs ( 42 1≤≤k) of \n180°-DWs [1], 10 MSCs ( 13 7≤≤k and 18 16 ≤≤k) of non 0°/180°- DWs and 42 MSCs ( \nk=2, 13 6≤≤k, 19 16 ≤≤k, k=22, 24, 26, 30, 32, 37, 39 and 64 43 ≤≤k) of 0°-DWs [4]. \nThe MSCs ( k=25, 28, 37-41, 52, 54, 61-63) with six-fold symmet ry axes (including inversion \naxes) do not realized in the cubic crystals [4]. \nThe unified co-ordinate system z y x O~~~ is chosen as [][ ]W12 zyx naaeee ,, ,, ~~~ −= where \nWn is the unit polar time-even vector along the DW pl ane normal [4]. For the 180°-DWs the \nvectors 1a and 2a are given early [1] as vectors 1τ and 2τ respectively. For the case of \n°≠180 2α the unit vector 1a coincides with the direction of the vector ()mnnm Δ−ΔWW (at \n0≠Δb and 0=Σb) or []Wna×2 (at 0=Δb or 0≠Σb) where 12mmm−=Δ , \n[]mnΔ ×=Δ Wb , []Σ Σ×=mnWb . Here the unit vector 2a coincides with the direction of \nvector ()Σ Σ−mnnmWW (at 0≠Σb) or []1Wan× (at 0≠Δb and 0=Σb) or else with an \narbitrary direction in the DW plane ( Wna⊥2 at 0 ==ΔΣbb ) where 2 1mmm +=Σ . The \nmutual orientation of the vectors 1m, 2m and Wn is determined by the parameters \n()Σ Σ=mnWa , ()mnΔ=Δ Wa , ()CW Camn= , Σb and Δb, where []21mmm ×=C . The mutual 3 \n \n orientation of the vectors 1m, 2m, Wn and Sn is determined by parameters ()Sana11=, \n()S 2 2ana=, ()SWnann= , []1S 1b an×= , []2S 2b an×= and []WSnb nn×= , where Sn \nis sample plane normal. \nThe MSC PG of restricted sample of crystal in paramagnetic ph ase could be defined \nas S P P GGG ∩=∞ where sample shape MSC SG is 1/mmm ′ ∞ for volumetric plate. MSCs of \nDWs in volumetric plate should satisfy the conditio n PkGG⊂. The MSCs of the all \npossible plane DWs in the arbitrary oriented plate of cubic crystals of hexoctahedral class \n(crystallographic point symmetry group m 3 m in the p aramagnetic phase [6]) are presented \nin table. Here symmetry axes are collinear with vec tors 1a and 2a and reflection planes are \nperpendicular to them. For MSCs with k =24, k =26, k =27, 29≤k≤36, 42 ≤k≤51, k =53, \n55≤k≤60 and k=64 only generative symmetry elements are represent ed. \nGeneral enumeration of MSCs of 0°-DWs contains MSCs with k=2, 13 6≤≤k, \n19 16 ≤≤k, k=22, 24, 26, 30, 32, 51 43 ≤≤k, k=53, 60 55 ≤≤k and k=64. The 60 0- and \n120°-DWs are represented by MSCs with k=10, 16, 18 and k=11, 13, 16 respectively. The \nMSCs of the 70.5 0-, 90 0- (both for <100> and <110> like easy magnetization axis [5]) and \n109.5 0-DWs are the MSCs with 7< k<13, 16< k<18. The general list of 180°-DWs includes \nMSCs with 42 1≤≤k except for k=25, 28, 37-41. \n \n \n \n \n \n \n \n 4 \n \n 3. Conclusions \nThe complete collection of ( nml )-plates with all possible orientations includes th e full \nlist of MSCs of α2 -DWs in cubic m 3 m crystal. For separate ( nml )-plates with fixed \ncombination of Miller indexes this list is limited. Such limitation depends on plate orientation. \nIt is minimal and maximal for the samples with high -symmetry (such as (100)-, (110)- or \n(111)-plates) and low-symmetry (the ( nml )-plates, where indexes are non-zero and have \ndifferent absolute values) developed surface respec tively. Maximal quantity of MSCs of α2-\nDWs is for (100)-plates. The MSC with k=16 is the MSC of all above-mentioned α2-DWs \nin arbitrary oriented plate of cubic m 3 mcrystal. \n \nReferences \n[1] V. Baryakhtar, V. Lvov, D. Yablonsky, JETP 87 , 1863 (1984) \n[2] V. Baryakhtar, E. Krotenko, D. Yablonsky, JETP 91 , 921(1986) \n[3] R. Vakhitov, A Yumaguzin, J. Magn. Magn. Mater. 215-216, 52 (2000) \n[4] B. M. Tanygin, O. V. Tychko, Physica B: Condensed Matter. Article in Press. \n[5] A Hubert and R. Shafer, Magnetic Domains. The Analysis of Magnetic \nMicrostructures , Springer, Berlin 1998 \n[6] L. Shuvalov, Modern Crystallography IV : Phys. Prop. Cryst., Springer, Berlin 1988 \n \n \n \n \n \n \n \n 5 \n \n Table. MSCs of the plane α2-DWs in plates of the cubic m3m crystal. \n \nk \n{nml }-sample 1m, 2m, Wn and \nSn: mutual \norientation**. 1m, 2m and Wn: \nmutual orientation Symmetry \n elements*** MSC \nsymbol \n1 {100}, {110} 0=nb or 1b= 0 0===ΔΣΣaba ()()1 , 12 , 2 , 2 , 112×n mmm \n2 {100}, {110} 0=nb or 1b= 0 0===ΣΔΔaba or \n0===ΔΣΣaba n2 , 2 , 2 , 12 1′′ mm′2′ \n3 {100}, {110} 0=nb or 1b= 0 0===ΔΣΣaba n2 , 2 , 2 , 112 mm 2 \n4 {nml }* 0=1a or 0=1b 0===ΔΣΣaba 1, '1,1' 2,12 /m 2′ \n5 {nml }* 0=na or 0=nb 0===ΔΣΣaba 1, '1,n' 2 ,n2 /m 2′ \n6 {nml }* 0a2=or 0b2= 0===ΔΣ Caaa \n22 , 1 m \n7 {100}, {110} 0=nb or 1b= 0 0==ΔΣaa 1, 12′,22,n2′ 22′2′ \n8 {nml }* 0=na or 0=nb 0==ΔΣaa 1, n2′ 2′ \n9 {100}, {110} 0=nb or 1b= 0 0===ΣΔbaaC 1, 12′,22′,n2 mm′2′ \n10 {nml }* 0=1aor 0=1b 0=Δa 1, 12′ 2′ \n11 {nml }* 0=na or 0=nb 0===ΣΔbaaC 1, n2 m \n12 {nml }* 0=1aor 0=1b 0=Ca 1, 12′ m′ \n13 {nml }* 0a2=or 0b2= 0=Σa 1, 22 2 \n14 {nml }* 0a2=or 0b2= 0==ΣΣba 1, 1,22′,22′ m / 2′′ \n15 Arbitrary Arbitrary 0==ΣΣba 1, '1 '1 \n16 Arbitrary Arbitrary Arbitrary 1 1 \n17 {100}, {110} 0=nb or 1b= 0 0===ΔΣbaaC 1, 12′,22,n2′ m′m′2 \n18 {nml }* 0=na or 0=nb 0===ΔΣbaaC 1, n2′ m′ \n19 {nml }* 0=na or 0=nb 0==ΣΔbb 1, n2 2 \n20 {nml }* 0=na or 0=nb 0 ===ΔΣΣbba 1, '1,n2,n2′ 2/m \n21 {100}, {110} 0=nb or 1b= 0 0===ΔΣΣbba 1, 12 ,22 ,n2 222 \n22 {100}, {110} 0=nb or 1b= 0 0==ΣΔbb 1, 12′,22′,n2 m′m′2 \n23 {100}, {110} 0=nb or 1b= 0 0===ΔΣΣbba ( )()' 1 , 12 , 2 , 2 , 121×n mmm'′′ \n24 {111} 0=nb 0==ΣΔbb n3 3 \n26 {111} 0=nb 0==ΣΔbb 12 , 3′n m 3′ \n27 {111} 0=nb 0===ΔΣΣbba 12 , 3n 32 \n29 {111} 0=nb 0===ΔΣΣbba 12 , ' 3′n m ' 3′ \n30 {100} 0=nb 0==ΣΔbb n4 4 \n31 {100} 0=nb 0 ===ΔΣΣbba nn' 2 , 4 m / 4′ \n32 {100} 0=nb 0==ΣΔbb 12 , 4′n mm4′′ \n33 {100} 0=nb 0===ΔΣΣbba 12 , 4n 422 \n34 {100} 0=nb 0===ΔΣΣbba nn' 2 , 2 , 41′ mm/m'4 ′′\n35 {100} 0=nb 0===ΔΣΣbba n' 4 '4 \n36 {100} 0=nb 0===ΔΣΣbba 12 , ' 4n m2'4′ \n42 {111} 0=nb 0 ===ΔΣΣbba n' 3 ' 3 6 \n \n Table. MSCs of the plane α2-DWs in plates of the cubic m3m crystal (continue). \n \nk \n{nml }-sample 1m, 2m, Wn and \nSn: mutual \norientation. 1m, 2m and Wn: \nmutual orientation Symmetry \nelements MSC \nsymbol \n43 {100}, {110} 0=nb or 1b= 0 0===ΣΔΔaba ()()1 , 12 , 2 , 2 , 12 1×′′n mmm′′ \n44 {100}, {110} 0=nb or 1b= 0 0===ΣΔΔaba n2 , 2 , 2 , 12 1′′ mm′2′ \n45 {nml}* 0a2=or 02=b 0===ΣΔΔaba 1, 1,22,22 2/m \n46 {nml }* 0=na or 0=nb 0===ΣΔΔaba 1, 1,n2′,n2′ m / 2′′ \n47 {nml }* 0=1aor 0=1b 0==ΔΔba 1, 1,12′,12′ m / 2′′ \n48 Arbitrary Arbitrary 0==ΔΔba 1, 1 1 \n49 {nml }* 0=na or 0=nb 0===ΣΔΔbba 1, 1,n2,n2 2/m \n50 {100}, {110} 0=nb or 1b= 0 0===ΣΔΔbba 1, 12′,22′,n2 22′2′ \n51 {100}, {110} 0=nb or 1b= 0 0===ΣΔΔbba ()()1 , 12 , 2 , 2 , 12 1×′ ′n mmm′′ \n53 {111} 0=nb 0===ΣΔΔbba 12 , 3′n 2 3′ \n55 {111} 0=nb 0===ΣΔΔbba 12 , 3′n m 3′ \n56 {100} 0=nb 0===ΣΔΔbba nn2 , 4 4/m \n57 {100} 0=nb 0===ΣΔΔbba 12 , 4′n 224′′ \n58 {100} 0=nb 0===ΣΔΔbba nn2 , 2 , 41′ mm/m 4′′ \n59 {100} 0=nb 0===ΣΔΔbba n4 4 \n60 {100} 0=nb 0===ΣΔΔbba 12 , 4′n m24′ ′ \n64 {111} 0=nb 0===ΣΔΔbba n3 3 \n \n* ( nml )-plates with arbitrary Miller indexes except non z ero values n≠m≠l≠n \n** At ()()02 1 = =ananW W \n *** The possible symmetry elements are rotations around two-fold symmetry axes n2 , n2′ or \n12, 12′ or else 22, 22′ that are collinear with the unit vectors Wn or 1a or else 2a, \nrespectively, reflections in planes n2, n2′ or 12 , 12′ or else 22 , 22′ that are normal to the \nabove mentioned vectors, respectively, rotations ar ound three-, four-fold symmetry axes n3 , \nn4 that are collinear with the vector Wn, rotations around three-, four-fold inversion \nsymmetry axes n3 ,n3′,n4 ,n4′ that are collinear with the vector Wn, inversion in the symmetry \ncenter 1, '1 and identity 1. Here an accent at symmetry element s means a simultaneous use \nof the time reversal operation [6]. \n " }, { "title": "1207.6039v2.High_cooperativity_in_coupled_microwave_resonator_ferrimagnetic_insulator_hybrids.pdf", "content": "High cooperativity in coupled microwave resonator ferrimagnetic insulator hybrids\nHans Huebl,1,\u0003Christoph W. Zollitsch,1Johannes Lotze,1Fredrik Hocke,1Moritz\nGreifenstein,1Achim Marx,1Rudolf Gross,1, 2and Sebastian T. B. Goennenwein1\n1Walther-Mei\u0019ner-Institut, Bayerische Akademie der Wissenschaften, Garching, Germany\n2Physik-Department, Technische Universit at M unchen, Garching, Germany\n(Dated: August 30, 2013)\nWe report the observation of strong coupling between the exchange-coupled spins in gallium-\ndoped yttrium iron garnet and a superconducting coplanar microwave resonator made from Nb. The\nmeasured coupling rate of 450 MHz is proportional to the square-root of the number of exchange-\ncoupled spins and well exceeds the loss rate of 50 MHz of the spin system. This demonstrates\nthat exchange coupled systems are suitable for cavity quantum electrodynamics experiments, while\nallowing high integration densities due to their spin densities of the order of one Bohr magneton per\natom. Our results furthermore show, that experiments with multiple exchange-coupled spin systems\ninteracting via a single resonator are within reach.\nPACS numbers: 85.25.-j, 85.70.Ge, 42.50.Pq, 76.30.+g\nKeywords: ferromagnetic resonance, ferrimagnetic resonance, strong coupling, cavity quantum electrody-\nnamics, YIG, yttrium iron garnet, low temperatures, microwave spectroscopy\nThe study of the interaction of matter and light on the\nquantum level is at the core of solid state quantum infor-\nmation systems. Strong [1, 2] and ultra-strong coupling\n[3] has been achieved, allowing for the coherent transfer\nof quantum information. For the practical implementa-\ntions of quantum information systems, the use of hybrid\nsystems has been suggested. In such hybrids, natural\nmicroscopic systems (atoms, molecules, electron spins,\nand nuclear spins) are coupled with arti\fcial meso-scale\nstructures such as superconducting quantum circuits by\nmeans of microwave photons [4{6]. Whereas the former\nhave long coherence times due to su\u000ecient decoupling\nfrom environmental noise, the latter allow for fast qubit\ngates due to strong coupling to electromagnetic \felds [7].\nEnsembles of electron spins as quantum memories [8, 9]\nseem promising and their coupling to superconducting\nresonators [10{16] and \rux qubits [17] has been studied\nrecently. Although the coupling strength gof an indi-\nvidual spin to the electromagnetic mode of a supercon-\nducting microwave resonator is small (typically 10 Hz),\nthe coupling of an ensemble of Nspins is enhanced by\na factor ofp\nN[18, 19]. In this way, strong coupling\nge\u000b=gp\nN\u001d\u0014;\rcan be realized, where \u0014and\rare the\nloss rates of the resonator and spin system, respectively.\nWith loss rates in the order of MHz, typically 1012spins\nare needed to reach the strong coupling regime. Until\ntoday, mostly paramagnetic systems consisting of ensem-\nbles of noninteracting spins have been studied. The co-\nherent coupling of microwave resonators to ferromagnetic\nsystems with strongly exchange coupled spins remains\nto be explored. Soykal and Flatt\u0013 e [20, 21] theoretically\ndiscussed the strong coupling of photonic and magnetic\nmodes in exchange locked ferromagnetic systems. Two\nparticular advantages of ferromagnetic systems are (i)\ntheir higher spin density, such that for the same number\nNof spins, their volume can be reduced considerablycompared to dilute paramagnetic systems, and (ii) the\nfact that below the magnetic ordering (Curie) tempera-\nture the system essentially is fully polarized, in contrast\nto the thermal polarization in uncoupled spin ensembles.\nThis should allow to couple multiple spin ensembles to\nthe same microwave resonator, e.g. for realizing an ad-\njustable coupling between the magnetic subsystems or for\nthe exchange of individual quanta between them.\nIn this letter, we investigate the coupling between the\nelectromagnetic modes of a superconducting coplanar\nwaveguide microwave resonator and the magnetic modes\nof the exchange-locked ferrimagnet yttrium iron garnet\n(Y3Fe5O12or YIG) doped with gallium (YIG:Ga). We\nmeasure a coupling rate of ge\u000b=2\u0019= 450 MHz exceed-\ning both the spin relaxation rate \r=2\u0019= 50 MHz and\nthe resonator decay rate \u0014=2\u0019= 3 MHz. That is, we\nobserve strong coupling. The measured e\u000bective cou-\npling strength follows ge\u000b=gp\nN, where the number\nof spins interacting with the resonator is estimated from\nthe sample geometry. Furthermore, the measured re-\nlaxation rate of the spin system is fully consistent with\nthe natural linewidth of YIG:Ga obtained from ferromag-\nnetic resonance (FMR) measurements [22]. Pure YIG is\none of the prime candidates for studying strong coupling\nbetween exchange locked spins and the electromagnetic\nmodes of a microwave resonator, in particular because\nof its very small FMR linewidth of \u001910\u0016T at 4 K and\n!=2\u0019= 9:3 GHz [23]. This narrow linewidth corresponds\nto aT2time in the order of microseconds [24]. Since\nhigh quality YIG thin \flms can be prepared on various\nsubstrates (gadolinium gallium garnet [25, 26], Si and\nGaAs [27]) and doped with rare earth elements in order\nto adjust the FMR linewidth in a controlled way, YIG\nseems an ideal material for ferromagnet based quantum\nhybrids.\nAs pointed out by Soykal and Flatt\u0013 e [20, 21], inarXiv:1207.6039v2 [quant-ph] 29 Aug 20132\nA\nYIG:GaB C\nT=50 mK/uni03BC0HT=0.7 KT=4.2 K\n3dB\n1dB3dB3dB\ncable cablecable cable3dBT=293 Kvector network analyzercable cable\nFIG. 1. Schematic of the experimental setup. The (purple)\ngallium doped YIG sample is cemented on top of one of the\nniobium microwave resonators which are arranged to allow\nfor multiplexing. Experiments are performed at millikelvin\ntemperatures in transmission by vector network analysis in a\nsuperconducting solenoid magnet.\na macrospin approximation the Hamiltonian for the\nferromagnet-resonator system can be expressed as\nH= \u0016h!raya+gs\u0016BBe\u000b\nzSz+ \u0016hg(aS++ayS\u0000):(1)\nHere,ayandaare the photon creation/annihilation oper-\nators,!rthe resonator frequency, gsthe electron g-factor,\n\u0016Bthe Bohr magneton, and Be\u000b\nzthe magnetic \feld [28].\nThe macrospin operator S= (S+^e\u0000\u0000S\u0000^e+)=p\n2 +Sz^z\nwith ^e\u0006=\u0007(^x\u0006{^y)=p\n2 is expressed in terms of the spin\nlowering and raising operators\nS\u0006\f\f\f\fN\n2;m\u001d\n=s\u0012N\n2\u0007m\u0013\u0012N\n2\u0006m+ 1\u0013\f\f\f\fN\n2;m\u00061\u001d\n:\n(2)\nHere,j`;miare the eigenstates of the macrospin and we\nhave assumed that the macrospin state is \fxed at its\nmaximal value `=N=2. In the Dicke model [29] of Nin-\ndependent paramagnetic spins this would correspond to\na fully excited spin system with no photons in the cavity.\nIn contrast to the Dicke model, for our macrospin model\nstates with ` < N= 2 are not accessible. Due to strong\nexchange coupling, states with ` < N= 2 are separated\nin energy and require the excitation of magnons. We\nnote that the coupling between the photonic and mag-\nnetic system is a magnetic dipole transition and that the\nHamiltonian (1) conserves the total excitation number\nZ=n+m, wherenis the photon number in the cavity\nandjmj\u0014`=N=2 the magnetic quantum number. As-\nsuming that Sis antiparallel to Bzandn= 0 initially,\nwe haveZ=N=2 and, hence, can index the basis states\njn;miof the resonator-spin systems either by the photon\nnumbern(jn;N\n2\u0000ni) or the magnetic quantum number\nm(jN\n2\u0000m;mi). Evidently, these basis states are similar\nto those of the Dicke model [29] for a paramagnetic en-semble ofNnoninteracting spins coupled to a resonator,\nwith`=N=2 taking the role of the cooperation number.\nDue to the analogy with the Dicke model, we expect\nthat the coupling strength of the ferromagnet-resonator\nsystem is given by the e\u000bective coupling strength ge\u000b=\ngp\nNof a paramagnet-resonator system, where g=\ngs\u0016B\n2\u0016hB1;0is the coupling rate of an individual spin with\nthe magnetic quantum number m= 1=2 to the resonator\n[29]. It is determined by the magnetic component of the\nrf vacuum \feld B1;0=p\n\u00160\u0016h!r=2Vmwhich depends on\nthe resonance frequency of the microwave resonator !r\nand its mode volume Vm[30]. Hence, for a given res-\nonance frequency, ge\u000b=gs\u0016B\n2\u0016hp\n\u00160\u001a\u0016h!rV=2Vmdepends\nonly on the spin density \u001a=N=V and the \flling factor\nV=V m, whereVdenotes the volume of the resonator \feld\nmode \flled with the spin system. Non-interacting spin\nensembles like paramagnetic centers in semiconductors or\ninsulators typically have a spin density \u001ain the order of\n1015\u0014\u001a\u00141018cm\u00003[10, 11, 13]. For !r=(2\u0019)\u00195 GHz,\nthis results in coupling rates in the order of 10 MHz as-\nsumingV=V m'1. In contrast, exchange coupled sys-\ntems naturally have a spin density in the order of one\nper atom (e.g. Fe, Ni, Co) or in the case of YIG 40 per\nunit cell (unit cell volume - 1 :8956 nm3), corresponding to\na spin density of 2 \u00021022cm\u00003[31]. Due to the increase\nof at least four orders of magnitude in the spin density\nwe expect a two orders larger coupling strength in ex-\nchange coupled systems as compared to non-interacting\nspins. Therefore, the exchange coupled system sample\nvolume can be reduced by a factor of 104while keeping\nthe coupling rate constant, enabling a higher integration\ndensity.\nIn our experiments, we study the coupling between the\nexchange locked system YIG:Ga and a superconducting\nNb resonator. The resonator structure is patterned into\na 100 nm thick Nb \flm deposited onto an intrinsic sili-\ncon substrate using optical lithography and reactive ion\netching [30]. Figure 1 shows the layout of the microwave\ncircuitry consisting of an input line, three resonators with\nresonance frequencies of fA= 5:90 GHz,fB= 5:53 GHz,\nandfC= 5:30 GHz, and an output line. This con\fgura-\ntion allows to compare loaded and unloaded microwave\nresonators on the same chip. A 2 \u00020:5\u00020:7mm3sized\n(length\u0002width\u0002thickness) commercial YIG:Ga crystal\nis cemented onto resonator Awith the highest microwave\nfrequencyfA= 5:90 GHz. The number of spins interact-\ning with the resonator is roughly estimated from the over-\nlap of the YIG:Ga crystal with the meandering coplanar\nwaveguide with a center conductor width of 6 \u0016m and a\ngap of 12 \u0016m. With the overlap length of 2 :5 mm and\nassuming that the vertical extension of the microwave\n\feld into the YIG crystal is about 30 \u0016m, the total num-\nber of spins coupled to the resonator is estimated to\nN\u00194:5\u00021016. To preserve the superconducting state\nof the microwave resonators, the surface of the chip is\ncarefully aligned in parallel to the applied magnetic \feld3\nA\nB\nC\nDFrequency (GHz)\nMagnetic Field B (mT)\n|S21|2 (10-5)1.6\n1.4\n1.2\n1.0\n0.8\n0.6\n0.4\n0.2\n06.5\n4.04.55.05.56.0\n0 200 100 300 0 200 100 300a b\next\nz\nFIG. 2. Transmission spectrum of the setup including the\nYIG-microwave resonator hybrid as a function of the applied\nmagnetic \feld Bext\nz, taken at T= 50 mK [33]. The copla-\nnar waveguide resonators ( A-C) show a slightly decreasing\nresonance frequency with increasing in-plane magnetic \feld.\nAdditionally, resonator Ashows a pronounced avoided cross-\ning at 170 mT, where the resonator frequency fAmatches the\nFMR frequency !FMR=(2\u0019). Resonance Dis a parasitic mode\npresent in the sample box. Panel a) shows the raw (uncali-\nbrated) transmission data as measured. Panel b) shows the\nsame data again superimposed with a \ft according to eq.(3)\nplotted as red line.\nBzgenerated by a superconducting solenoid. The mi-\ncrowave transmission experiments are performed at the\nbase temperature of a dilution refrigerator of 50 mK us-\ning a commercial vector network analyzer. To thermally\nanchor the center conductor of the microwave input and\noutput lines, attenuators are used at the 4 K, the still and\nthe mixing chamber stages (cf. Fig. 1). Considering only\nthe attenuators, we estimate a microwave \feld tempera-\nture of about 70 K (or 290 thermally excited photons on\naverage) in the resonator. [32]\nFigure 2 shows the microwave transmission jS21j2raw\ndata as a function of frequency and applied magnetic\n\feld [33]. In the spectrum at Bz= 0, four transmis-\nsion peaks are visible corresponding to the resonance fre-\nquencies of the coplanar microwave resonators A,B, and\nC. The broad feature labeled Dstems from a parasitic\nmode of the metallic microwave box in which the sam-\nple is mounted. As expected, the resonators BandC\nshow only a weak magnetic \feld dependence, because\nthey are not interacting with the YIG:Ga crystal due to\nthe absence of physical overlap (cf. Figs. 1 and 2). On\nthe contrary, resonator Aand the box mode Dcouple\nto the YIG:Ga. While mode Dallows us to probe the\nferromagnetic resonance independently of the strongly\ncoupled mode ( A) and to determine the FMR disper-\nsion relation, resonator Ashows a distinct anticrossing\natBext\nz(\u0001 = 0) = BFMR = 170 mT where the FMR\ndispersion relation \u0016 h!FMR =gs\u0016BBe\u000b\nz[34] is degeneratewith the resonator \u0016 h!r.\nTo derive the e\u000bective coupling rate ge\u000bfrom the mea-\nsured data, we simplify the discussion of (1), by modeling\nthe system as two coupled harmonic oscillators. The dis-\npersion of the resonance frequency is then given by [35]\n!=!r+\u0001\n2\u00061\n2q\n\u00012+ 4g2\ne\u000b: (3)\nHere, \u0001 = !FMR\u0000!r=gs\u0016B(Bext\nz\u0000BFMR)=\u0016his\nthe \feld dependent detuning between the resonator fre-\nquency!r=(2\u0019) =fAand the \feld dependent FMR fre-\nquency!FMR=(2\u0019). The experimental data agree very\nwell with this model prediction. Fitting the data yields\nge\u000b= 450\u000620 MHz and BFMR = 170\u00065 mT, and the\ng-factor of the ferrimagnetic resonance gs= 2:17\u00060:05\n(red line in Fig.2). Here, BFMR is reduced with respect\nto the bare electron spin resonance \feld of 194 mT due to\nthe presence of an anisotropy \feld Ba= 24 mT [28],[34].\nAdditionally, the experimentally observed gsis not ex-\nactly identical to the literature value for pure YIG at 2 K\nofgs;lit= 1:99 [36] or YIG:Ga at 5 K of gs;lit= 2:1[22].\nThe observed di\u000berence might be due to the higher Ga\nconcentration compared to Ref. [22] or the lower tempera-\nture. Note, that the g-factor and the magnetic anisotropy\nof YIG and doped YIG is not well established and re-\nquires further investigations. For our resonator we esti-\nmateg=(2\u0019)'5 Hz [30] and N'4\u00021016. With these\nnumbers we expect ge\u000b=(2\u0019) = (g=2\u0019)p\nN'1 GHz cor-\nroborating our experimental result within a factor of two\ndespite of the rough estimate for N.\nTo determine the relaxation rate \rof the spin system\nwe analyze the evolution of the linewidth of the resonator\nmodeAas a function of the magnetic \feld Bzby \ftting\na Lorentzian lineshape in the frequency domain for ev-\nery measured magnetic \feld magnitude [16]. Figure 3(a)\nshows the resonance frequency obtained from such a \ft\nas red crosses superimposed on the color-coded dataset.\nIn Fig. 3(b) the corresponding (FWHM) linewidth (red\ncrosses) is shown. At low magnetic \felds, the resonator\nmodeAis essentially decoupled from the spin system,\nsuch that the measured linewidth is given by \u0014. Closer to\nthe ferromagnetic resonance, the linewidth of the system\nis given by the combined relaxation rate of the spin sys-\ntem and the microwave resonator leading to an increase\nin the observed linewidth.\nTo quantify the coupling and loss rates in our system,\nwe use a standard input-output formalism [10, 16, 19,\n37]. Within this framework the transmission amplitude\nof microwaves from the input to the output port of the\nmicrowave resonator is given by\nS21=\u0014c\n{(!\u0000!r)\u0000(\u0014c+\u0014i) +jgeffj2\n{(!\u0000!FMR)\u0000\r=2:(4)\nHere,!=2\u0019is the frequency of the microwave probe tone,\n\u0014cis the external coupling rate between the microwave4\nMagnetic Field B (mT)ext\nz250 200 150 100 50 298Frequency\n(GHz)Frequency\n(GHz)\n|S21|2 (10-6)2.5\n2.0\n1.5\n1.0\n0.5\n0Peak Width\n(MHz)6.2\n5.86.06.2 20 40 60 805.86.0a\ncb\nDataSimulation\nFIG. 3. Analysis of the resonance frequency and linewidth as\na function of the external magnetic \feld Bext\nz. Panel (a) shows\nthe resonator transmission jS21j2(same data as Fig. 2) as a\nfunction of frequency and applied magnetic \feld [33]. The red\ncrosses mark the resonance frequency determined by \ftting a\nLorentzian lineshape to the data at constant Bext\nz. The red\ncrosses in panel (b) show the extracted FWHM linewidths\ncorresponding to 2 \r=(2\u0019) and 2\u0014=(2\u0019). In addition, the blue\ncircles show the linewidths obtained from the numerical sim-\nulation of the transmission spectra plotted in panel (c). The\nsimulation is based on the input-output formalism resulting\nfrom eq.(4) [10, 16, 19, 37].\nresonator and the feed line, and \u0014isummarizes the in-\ntrinsic loss rate of the microwave resonator. In our case\nwe have\u0014i\u001d\u0014c, resulting in a total microwave res-\nonator relaxation rate \u0014'\u0014i(cf. [34]). Fig. 3(c) shows\nthe calculated transmission using ge\u000b=2\u0019= 450 MHz,\n\r=2\u0019= 50 MHz, and \u0014=2\u0019= 3 MHz. Evidently, all\nfeatures of the experimental data of Fig. 3(a) are nicely\nreproduced. Moreover, the two transmission peaks ex-\npected atBFMR = 170 mT cannot be resolved due to the\nlimited signal to noise ratio in the experimental data.\nAdditonally, we can analyze the simulation data shown\nin Fig. 3(c) in the same way as the experimental data\nin Fig. 3(a). The result is shown by the blue circles in\nFig. 3(b), where in contrast to the experimental data,\nthe modeled data is noise-free allowing to predict the\nlinewidth for all magnetic \feld values. The good agree-\nment between experimental and simulation data again\ndemonstrates that the parameters chosen in the simula-\ntion well reproduce the experimental situation. In sum-\nmary, our analysis shows that ge\u000b\u001d\u0014;\rwith a cooper-\nativityC=g2\ne\u000b=\u0014\r'1350. That is, the strong coupling\nregime has been reached for the ferrimagnet-resonator\nsystem [38].\nNext, we compare the experimentally determined re-\nlaxation rate \rwith the temperature dependence of\nthe FMR linewidth measured at 9 :43 GHz. Typically,\nYIG:Ga exhibits signi\fcantly larger damping (larger\nlinewidth) than pure YIG. Rachford et al. [22] report\nlinewidths for YIG:Ga that decrease from 1 mT at 4 :2 Kto 0:1 mT at room temperature, corresponding to 28 MHz\nand 2:8 MHz, respectively. Our sample has a higher\nGa-doping concentration [39], and thus larger linewidth,\nwhich coroborates the relaxation rate \r=2\u0019= 50 MHz\nwe measured at millikelvin temperatures. However, note\nthat little is known about the damping in YIG for T\u0014\n2 K. According to Sparks and Kittel [40] the limiting re-\nlaxation mechanism is spin-lattice coupling, expected to\nbe well below 1 MHz for YIG. This calls for further ex-\nperiments in this temperature regime to verify the pro-\nposed relaxation mechanism and to eludicate the maxi-\nmum spin coherence time achievable in YIG for T\u00142 K.\nNote also, that relaxation mechanisms in ferromagnets\nare fundamentally di\u000berent from relaxation mechanisms\nin diluted paramagnetic systems. In the latter, spin-spin\ninteractions can cause dephasing and decoherence [41].\nIn the former, it is possible to simultaneously have high\nspin density and low damping.\nFinally, we \fnd that ge\u000bis independent of the mi-\ncrowave power from 10 fW to 10 nW. This is expected,\nbecause here the number of excitations (photons in the\nmicrowave resonator) is much smaller than the number\nof spinsN\u00191016[14]. Only if the number of excita-\ntions (photons) becomes comparable to or exceeds N, a\nquenching of the observed anticrossing is expected.\nIn conclusion, we experimentally observed strong cou-\npling between a superconducting microwave resonator\nand a gallium doped yttrium iron garnet ferrimagnet.\nThe e\u000bective coupling rate of 450 MHz reaches 8% of the\nresonator frequency !r=(2\u0019) and by far exceeds the re-\nlaxation rate \r=(2\u0019) = 50 MHz of the spin system at\nabout 50 mK, which is rather large owing to the gallium\ndoping. Considering the much smaller linewidth in pure\nYIG, even higher cooperativities can be anticipated. Our\nresults establish that exchange coupled spin systems in-\ndeed can be used for cavity quantum electrodynamics.\nFurthermore, the large coupling rates achievable in ex-\nchange coupled systems allow to place more than one\nmagnetic system in the microwave resonator. This should\nallow to study e.g. the exchange of magnetic excitations\nvia a cavity bus similar to the approaches pursued in the\n\feld of cavity QED [7].\nWe acknowledge technical support by M. Opel. This\nwork is supported by the German Research Foundation\nthrough SFB 631 and the German Excellence Initiative\nvia the \\Nanosystems Initiative Munich\" (NIM).\n\u0003corresponding author huebl@wmi.badw.de\n[1] A. Wallra\u000b, D. I. Schuster, A. Blais, L. Frunzio, R.-S.\nHuang, J. Majer, S. Kumar, S. M. Girvin, and R. J.\nSchoelkopf, Nature 431, 162 (2004).\n[2] R. J. Schoelkopf and S. M. Girvin, Nature 451, 664\n(2008).\n[3] T. Niemczyk, F. Deppe, H. Huebl, E. P. Menzel,5\nF. Hocke, M. J. Schwarz, J. J. Garcia-Ripoll, D. Zueco,\nT. H ummer, E. Solano, A. Marx, and R. Gross, Nat.\nPhys. 6, 772 (2010).\n[4] A. Andr\u0013 e, D. DeMille, J. M. Doyle, M. D. Lukin, S. E.\nMaxwell, P. Rabl, R. J. Schoelkopf, and P. Zoller, Nat.\nPhys. 2, 636 (2006).\n[5] J. Verd\u0013 u, H. Zoubi, C. Koller, J. Majer, H. Ritsch, and\nJ. Schmiedmayer, Phys. Rev. Lett. 103, 043603 (2009).\n[6] M. Wallquist, K. Hammerer, P. Rabl, M. Lukin, and\nP. Zoller, Phys. Scr. T137 , 014001 (2009).\n[7] L. DiCarlo, J. M. Chow, J. M. Gambetta, L. S. Bishop,\nB. R. Johnson, D. I. Schuster, J. Majer, A. Blais, L. Frun-\nzio, S. M. Girvin, and R. J. Schoelkopf, Nature 460, 240\n(2009).\n[8] J. H. Wesenberg, A. Ardavan, G. A. D. Briggs, J. J. L.\nMorton, R. J. Schoelkopf, D. I. Schuster, and K. Molmer,\nPhys. Rev. Lett. 103, 070502 (2009).\n[9] A. Imamoglu, Phys. Rev. Lett. 102, 083602 (2009).\n[10] D. I. Schuster, A. P. Sears, E. Ginossar, L. DiCarlo,\nL. Frunzio, J. J. L. Morton, H. Wu, G. A. D. Briggs,\nB. B. Buckley, D. D. Awschalom, and R. J. Schoelkopf,\nPhys. Rev. Lett. 105, 140501 (2010).\n[11] Y. Kubo, F. R. Ong, P. Bertet, D. Vion, V. Jacques,\nD. Zheng, A. Dreau, J. F. Roch, A. Au\u000beves, F. Jelezko,\nJ. Wrachtrup, M. F. Barthe, P. Bergonzo, and D. Esteve,\nPhys. Rev. Lett. 105, 140502 (2010).\n[12] Y. Kubo, C. Grezes, A. Dewes, T. Umeda, J. Isoya,\nH. Sumiya, N. Morishita, H. Abe, S. Onoda, T. Ohshima,\nV. Jacques, A. Dr\u0013 eau, J.-F. Roch, I. Diniz, A. Au\u000beves,\nD. Vion, D. Esteve, and P. Bertet, Phys. Rev. Lett. 107,\n220501 (2011).\n[13] R. Ams uss, C. Koller, T. N obauer, S. Putz, S. Rot-\nter, K. Sandner, S. Schneider, M. Schramb ock, G. Stein-\nhauser, H. Ritsch, J. Schmiedmayer, and J. Majer, Phys.\nRev. Lett. 107, 060502 (2011).\n[14] I. Chiorescu, N. Groll, S. Bertaina, T. Mori, and\nS. Miyashita, Phys. Rev. B 82, 024413 (2010).\n[15] P. Bushev, A. K. Feofanov, H. Rotzinger, I. Protopopov,\nJ. H. Cole, C. M. Wilson, G. Fischer, A. Lukashenko,\nand A. V. Ustinov, Phys. Rev. B 84, 060501 (2011).\n[16] E. Abe, H. Wu, A. Ardavan, and J. J. L. Morton, Appl.\nPhys. Lett. 98, 251108 (2011).\n[17] X. Zhu, S. Saito, A. Kemp, K. Kakuyanagi, S. Kari-\nmoto, H. Nakano, W. J. Munro, Y. Tokura, M. S. Everitt,\nK. Nemoto, M. Kasu, N. Mizuochi, and K. Semba, Na-\nture478, 221 (2011).\n[18] M. G. Raizen, R. J. Thompson, R. J. Brecha, H. J. Kim-\nble, and H. J. Carmichael, Phys. Rev. Lett. 63, 240\n(1989).\n[19] D. F. Walls and G. J. Milburn, Quantum Optics, 1st ed.\n(Springer, Berlin, 1994).\n[20] O. O. Soykal and M. E. Flatt\u0013 e, Phys. Rev. B 82, 104413\n(2010).\n[21] O. O. Soykal and M. E. Flatt\u0013 e, Phys. Rev. Lett. 104,\n077202 (2010).\n[22] F. J. Rachford, M. Levy, R. M. Osgood, A. Kumar, and\nH. Bakhru, J. Appl. Phys. 87, 6253 (2000).\n[23] E. G. Spencer, R. C. LeCraw, and A. M. Clogston, Phys.\nRev. Lett. 3, 32 (1959).\n[24] D. E. Kaplan, Phys. Rev. Lett. 14, 254 (1965).\n[25] S. A. Manuilov, R. Fors, S. I. Khartsev, and A. M. Gr-\nishin, J. Appl. Phys. 105, 033917 (2009).\n[26] S. A. Manuilov and A. M. Grishin, J. Appl. Phys. 108,\n013902 (2010).[27] M. Levy, R. M. Osgood, A. Kumar, and H. Bakhru,\nAppl. Phys. Lett. 71, 2617 (1997).\n[28] The magnetic induction Be\u000b\nz=Bext\nz+Bahas two contri-\nbutions: i) the externally applied magnetic \feld Bext\nzand\nii) internal \felds accounting for the magnetic anisotropy\nBa. (cf. [34]).\n[29] R. H. Dicke, Phys. Rev. 93, 99 (1954).\n[30] T. Niemczyk, F. Deppe, M. Mariantoni, E. P. Menzel,\nE. Ho\u000bmann, G. Wild, L. Eggenstein, A. Marx, and\nR. Gross, Supercond. Sci. Tech. 22, 034009 (2009).\n[31] M. Gilleo and S. Geller, Phys. Rev. 110, 73 (1958).\n[32] This accounts for the \fxed attenuators and does not in-\nclude the lossy lines. For further details refer to [34].\n[33] Since we have not calibrated our setup at millikelvin tem-\nperatures, we show raw, uncalibrated, as-measured S21\ntransmission data in Figs. 2 and 3. In our opinion, this is\nthe most honest way of presenting the data in lack of a\nproper full calibration. .\n[34] \\Sublemental material,\".\n[35] S. Haroche and J. M. Raimond,\nExploring the Quantum: Atoms, Cavities and Photons\n(Oxford University Press, Oxford, 2006).\n[36] K. P. Belov, L. A. Malevskaya, and V. I. Sokolov, Soviet\nPhysics JETP 12, 1074 (1960).\n[37] A. A. Clerk, S. M. Girvin, F. Marquardt, and R. J.\nSchoelkopf, Rev. Mod. Phys. 82, 1155 (2010).\n[38] Our claims are based on the analysis of the lineshape\nand resonance frequency dependence as function of the\napplied magnetic \feld and do not rely on the absolute\ntransmission intensity. We therefore have not calibrated\nthe setup with respect to the amplitude information.\n[39]YIG:Ga Datasheet - www.ferrisphere.com.\n[40] M. Sparks and C. Kittel, Phys. Rev. Lett. 4, 232 (1960).\n[41] A. M. Tyryshkin, S. Tojo, J. J. L. Morton, H. Riemann,\nN. V. Abrosimov, P. Becker, H.-J. Pohl, T. Schenkel,\nM. L. W. Thewalt, K. M. Itoh, and S. A. Lyon, Nature\nMaterials 11, 143 (2011)." }, { "title": "1804.08719v1.Unidirectional_Loop_Metamaterials__ULM__as_Magnetless_Artificial_Ferrimagnetic_Materials__Principles_and_Applications.pdf", "content": "1\nUnidirectional Loop Metamaterials (ULM)\nas Magnetless Artificial Ferrimagnetic Materials:\nPrinciples and Applications\nToshiro Kodera, Senior Member, IEEE, and Christophe Caloz, Fellow, IEEE\nAbstract —This paper presents an overview of Unidirectional\nLoop Metamaterial (ULM) structures and applications. Mimick-\ning electron spin precession in ferrites using loops with unidi-\nrectional loads (typically transistors), the ULM exhibits all the\nfundamental properties of ferrite materials, and represents the\nonly existing magnetless ferrimagnetic medium . We present here\nan extended explanation of ULM physics and unified description\nof its component and system applications.\nIndex Terms —Unidirectional Loop Metamaterials (ULM),\nnonreciprocity, ferrimagnetic materials and ferrites, gyrotropy,\nFaraday rotation, metamaterials and metasurfaces, transistors,\nisolators, circulators, leaky-wave antennas.\nI. I NTRODUCTION\nOver the past decades, nonreciprocal components (isola-\ntors, circulators, nonreciprocal phase shifters, etc.) have been\nhave been almost exclusively implemented in ferrite tech-\nnology [1]–[8]. This has been the case in both microwaves\nand optics, despite distinct underlying physics, namely the\npurely magnetic effect (electron spin precession) in the former\ncase [3], [9], [10] and the magneto-optic effect (electron\ncyclotron orbiting) [11]–[13] in the latter case. However,\nferrite components suffer from the well-known issues high-\ncost, high-weight and incompatibility with integrated circuit\ntechnology, and magnetless nonreciprocity has therefore long\nbeen consider a holy grail in this area [14], [15].\nThere have been several attempts to develop magnetless\nnonreciprocal components, specifically 1) active circuits [16]–\n[21], and space-time [15] 2) modulated structures [22]–[27]\nand 3) switched structures [28] (both based on 1950ies para-\nmetric (e.g. [29], [30]) or commutated (e.g. [31]) microwave\nsystems). All have their specific features, as indicated in Tab. I.\nTABLE I\nCOMPARISON (TYPICAL AND RELATIVE TERMS )BETWEEN DIFFERENT\nMAGNETLESS NONRECIPROCITY TECHNOLOGIES PLUS FERRITE .\nmaterialPconsum. bias cost noise\nferrite yes zero magnet high N/A\nact. circ. no low DC low low\nswitched no med. RF high high\nmodulated no med. RF med. med.\nULM YES low DC low med\nWe introduced in 2011 [32] in a Unidirectional Loop\nMetamaterial (ULM) mimicking ferrites at microwaves and\nrepresenting the only artificial ferrite material, or metamaterial,\nexisting to date. This paper presents an overview of the ULM\nand its applications reported to date.\nT. Kodera is with the Department of Electrical Engineering, Meisei Univer-\nsity, Tokyo Japan (e-mail: toshiro.kodera@meisei-u.ac.jp). C. Caloz is with the\nDepartment of Electrical and Engineering, ´Ecole Polytechnique de Montr ´eal,\nMontr ´eal, QC, H2T 1J3 Canada.II. O PERATION PRINCIPLE\nA Unidirectional Loop Metamaterial (ULM) may be seen\nas a physicomimetic1artificial implementation of a ferrite in\nthe microwave regime . Its operation principle is thus based\nonmicroscopic unidirectionality , from which the macroscopic\ndescription is inferred upon averaging.\nA. Microscopic Description\nMicrowave magnetism in a ferrite is based on the precession\nof the magnetic dipole moments arising from unpaired electron\nspins about the axis of an externally applied static magnetic\nbias field, B0, as illustrated in Fig. 1(a), where B0k^ z. This is\na quantum-mechanical phenomenon, that is described by the\nLandau-Lifshitz-Gilbert equation [3], [10], [33]\ndm\ndt=\u0000\rm\u0002B0+\u000b\nMsm\u0002dm\ndt; (1)\nwhere mdenotes the magnetic dipole moment, \rthe gyromag-\nnetic ratio,Msthe saturation magnetization, and \u000bthe Gilbert\ndamping term. Equation (1) states that the time-variation rate\nofmdue to a transverse2RF magnetic field signal, HRF\nt\n(k^t;^t?^ z), is equal to the sum of the torque exerted by B0\nonm(directed along +^\u001e,^\u001e: azimuth angle), and a damping\nterm (directed along\u0000^\u0012,^\u0012: elevation angle) that reduces\nthe precession angle, , to zero along a circular-spherical\ntrajectory (conserved jmj) when the RF signal is suppressed\n(relaxation).\nClassically, magnetic dipole moments can be associated\nwith current loop sources , according to Amp `ere law. Decom-\nposing a ferrite magnetic moment, m, into its longitudinal\ncomponent, mz, and transverse component, m\u001a, as shown in\nFig. 1(a), one may thus invoke the effective current loops Imz\neff\nandim\u001a\neffassource models for the corresponding moments.\nAmong these currents, only im\u001a\neffmatters in terms of mag-\nnetism, since Imz\neff, as the source associated with HRF\nz, does\nnot induce any precession (Footnote 2). im\u001a\neffis thus the current\none has to mimic to devise an “artificial ferrite.” This current,\nas seen Fig. 1, has the form of a loop tangentially rotating on\nan imaginary cylinder of axis z.\n1The adjective “physicomimetic” is meant here, from etymology, as “mim-\nicking physics.”\n2The longitudinal ( z) component does not contribute to precession, and\nhence to magnetism. Indeed, since B0k^ z, thez-component of mproduced by\nHRF\nzwould lead to mRF\nz\u0002(B0+\u00160HRF^ z) = [mRF\nz(B0+\u00160HRF)](^ z\u0002^ z) =\n0, the only torque being produced by the transverse component ( HRF\nt,^t2xy-\nplane), mRF\nt\u0002(B0+\u00160HRF^t) = (mRF\ntB0)(^t\u0002^ z)6= 0. In the rest of the\ntext, we shall drop the superscript “RF,” without risk of ambiguity since mt,\nis exclusively produced by the RF signal.arXiv:1804.08719v1 [physics.app-ph] 16 Apr 20182\nFig. 1. “Physicomimetic” construction of the Unidirectional Loop Metama-\nterial (ULM) “meta-molecule” or particle. (a) Magnetic dipole precession,\narising from electron spinning in a ferrite material about the axis (here z) of an\nexternally applied static magnetic bias field, B0, with effective unidirectional\ncurrent loops Imz\neffandim\u001a\neff, and transverse radial rotating magnetic dipole\nmoment m\u001aassociated with im\u001a\neff. (b) ULM particle [32], typically (but not\nexclusively [34]) consisting of a pair of broadside-coupled transistor-loaded\nrings supporting antisymmetric current and unidirectional current wave (shown\nhere with exaggeratedly small wavelength for the sake of visibility), with\nresulting radial rotating magnetic dipole moment emulating that in (a).\nGiven its complexity, the current im\u001a\neffmay a priori seem\nimpossible to emulate. However, what fundamentally matters\nfor magnetism is not this current itself, but the moment m\u001a,\nfrom which magnetization will arise at the macroscopic level\n(Sec. II-B). This moment may be fortunately also produced\nby a pair of antisymmetric \u001e-oriented currents, rotating on the\nsame cylinder, which can be produced by a pair of conducting\nrings operating in their odd mode [35], as shown in Fig. 1(b).\nIf this ring-pair structure is loaded by a transistor [32], as\ndepicted in the figure, or includes another unidirectionality\nmechanism such as the injection of an azimuthal modula-\ntion [34], m\u001awillunidirectionally rotate about zwhen excited\nby an RF signal, and hence mimic the magnetic behavior of the\nelectron in Fig. 1(b). The structure in Fig. 1 constitutes thus\ntheunit-cell particle of the ULM at the microscopic level .\nOne may argue there there is a fundamental difference\nbetween the physical system in Fig. 1(a) and its presumed\nartificial emulation in Fig. 1(b): the ferrite material also\nsupports the longitudinal moment mzwhereas the ULM does\nnot include anything alike. However, as we have just seen\nabove, particularly in Footnote 2, mzdoes not contribute to\nthe magnetic response. It is therefore inessential and does thus\nnot need to be emulated. So, the particle in Fig. 1(b), with its\nmoment m\u001ais all that is needed for artificial magnetism !\nDoes this mean that the ULM particle includes no counter-\npart to the static alignment of the dipoles due to B0(and\nproducing mz) in the ferrite medium? In fact, there isa\ncounterpart, although mz= 0 in the ULM. The fundamental\noutcome of the static alignment of the dipoles along zin\nthe ferrite is the alignment of the relevant magnetic dipoles\nm\u001ain the plane perpendicular to ^ z(or within the xy-plane)\nacross the medium, for otherwise the m\u001a’s of the different\ndomains [3], [10] would macroscopically cancel out. Such an\norientation of the m\u001a’s perpendicularly to ^ zis essential to\nemulate magnetism. How is this provided in the ULM? Simply\nbyfixing the rings in a mechanical support , such as a substrate,\nas will be seen later. So, the counterpart of the ferrite static\nalignment of dipoles is simply mechanical orientation in the\nULM.B. Macroscopic Description\nSince it mimics the relevant magnetic operation of a ferrite\nat the microscopic level, the unit-cell particle in Fig. 1(b) must\nlead to the same response as bulk ferrite at the macroscopic\nlevel when repeated according to a subwavelength 3D lattice\nstructure so as to form a metamaterial as shown in Fig. 2(a).\nThe ULM in Fig. 2(a), just as a ferrite3, forms a 3D\narray of magnetic dipole moments, mi, whose average over a\nsubwavelength volume V,\nM=1\nVX\ni=1mi= \n1\nVX\ni=1m\u001a;i!\n^\u001a=M\u001a^\u001a (2)\ncorresponds to the density of magnetic dipole moments, or\nmagnetization , as the fundamental macroscopic quantity de-\nscribing the metamaterial4.\n(a) (b)\nFig. 2. ULM structures obtained by periodically repeating the unit cell with\nthe particle in Fig. 1. (a) Metamaterial (3D), described by the Polder volume\npermeability (3a). (b) Metasurface (2D metamaterial), described by a surface\npermeability [36].\nFrom this point, one may follow the same procedure as in\nferrites [8], [37] to obtain the Polder ULM permeability tensor\n\u0016=2\n4\u0016 j\u0014 0\n\u0000j\u0014 \u0016 0\n0 0\u00160;3\n5; (3a)\nwith\u0016=\u00160\u0012\n1 +!0!m\n!2\n0\u0000!2\u0013\nand\u0014=\u00160!!m\n!2\n0\u0000!2;(3b)\nwhere!0and!mare the ULM resonance frequency (or\nLarmor frequency) and effective saturation magnetization fre-\nquency , respectively5, that will be derived in the next section.\nAs in ferrites, the effect of loss can be accounted for by\nthe substitution !0 !0+j\u000b!, where\u000ba damping factor\nin (1) [8].\nSo, a ULM may really be seen as a an artificial ferrite\nmaterial producing magnet-less artificial magnetism . However,\nitsnonreciprocity is achieved from breaking time-reversal\n(TR) symmetry by a TR-odd current bias , originating in the\ntransistor (DC) biasing, instead of a TR-odd external magnetic\nfield [14].\nULMs have been implemented only in a 2D format so far.\nThe corresponding structure is shown in Fig. 2(b), and may\n3The difference is essentially quantitative : while in the ferrite p=\u0015< 10\u00006\n(p: molecular lattice constant), in the ULM p=\u0015\u00191=10\u00001=5(p:\nmetamaterial lattice constant or period), but homogeneization works in both\ncases.\n4Whereas in a ferrite, we have M=Ms+MRF= (Ms+MRF\nz)^ z+MRF\nt\u0019\nMs^ z+MRF\n\u001a^\u001a, whereMsis the saturation magnetization of material, in the\nULMMs= 0. We shall subsequently drop the superscript “RF” also in M\u001a.\n5In a ferrite,!0=\rB0and!m=\r\u00160Ms.3\nbe referred to as a Unilateral Loop Metasurface (ULMS) .\nSection IV will present ULMS Faraday rotation and Sec. V-A\nwill discuss related applications.\nIII. ULM P ARTICLE AND DESIGN\nULMs may be implemented in different manners. Figure 3\nshows a ULM particle implemented in the form of a microstrip\ntransistor-loaded single ring placed on PEC plane. Assuming\na distance much smaller than the wavelength between the\nring and the PEC plane, the structure is equivalent, by the\nimage principle, to the antisymmetric double-ring structure in\nFig. 1(b) [35].\nFig. 3. ULM particle microstrip implementation in the form of a transistor-\nloaded single ring on a PEC plane, supporting the odd effective current\ndistribution and unidirectional current wave as in Fig. 1(b). The transistor\nbiasing circuit is not shown here.\nThe transistor-loaded ULM particle in Fig. 3, or Fig. 1(b),\nis essentially a ring resonator , whose total electrical size is\ngiven by [38]\n\fmsa(2\u0019\u0000\u000bTR) +'TR= 2\u0019; \f ms=k0p\u000fe=!\ncp\u000fe;(4)\nwhere\fmsis the microstrip line wavenumber ( \u000fe: effective\nrelative permittivity), ais the average radius of the ring, \u000bTR\nis the geometrical angle subtending the transistor chip, and\n'TRis the phase shift across it. Solving Eq. (4) for !provides\nthe resonance frequency of the resonator, and hence the ULM\nresonance frequency ,\n!0=\f\f\f\f(2\u0019\u0000'TR)c\nap\u000fe(2\u0019\u0000\u000bTR)\f\f\f\f; (5)\nin (3). The parameter !min the same relations follows from\nthe mechanical orientation of the moments, as explained in\nSec. II-B: although we do not have here a saturation magne-\ntizationMsleading to the frequency parameter !m=\r\u00160Mm\nin the ferrite, we have have an equivalent phenomenological\nparameter!massociated with the orientation of the rings,\nwhich may be found by extraction, as will be seen in Sec. IV.\nNote that ULMs may be designed for multi-band operation\nand enhanced-bandwidth operation. The former, in contrast\nto ferrites that are restricted to a single ferromagnetic reso-\nnance!0=\rB06, can in principle accommodate multiple\nresonances by simply incorporating rings of different sizes,\nas illustrated in Fig. 4. The latter, in contrast to ferrite whose\nbandwidth is inversely proportional to loss due to causality, can\nbe achieved by leveraging overlapping coupled resonances [2].\n6This restriction can be somewhat overcome in a structured ferromagnetic\nstructure, such as a ferromagnetic nanowire membrane supporting a remanent\nbistable population of up and down magnetic dipole moments with corre-\nsponding resonances !\"\n0=\r\u00160H\"\n0and!#\n0=\r\u00160H#\n0[39], [40].\n(a) (b)Fig. 4. Unique magnetic properties of the ULM attainable by using multiple\nrings of resonance frequencies !0n, here two rings with resonance frequencies\n!01and!02. (a) Multiband operation using independent rings with separate\nresonance frequencies. (b) Enhanced bandwidth using coupled-resonant rings\nwith overlapping resonance frequencies.\nThe single-ring-PEC ULM implementation of Fig. 3 is\nideal for microstrip components [38] and reflective metasur-\nfaces [32]. However, for 3D ULM [Fig. 2(a)] structures and\ntransmissive metasurfaces (Sec. IV), a transparent version of\nthat ULM is required. This could theoretically be realized with\na pair of rings, as in Fig. 1(b), but may be more conveniently\nimplemented in the form of circular slots in a Coplanar Wave-\nGuide (CPW) type technology, as reported in [41].\nIV. F ARADAY ROTATION\nFaraday rotation is one of the most fundamental and useful\nproperties of magnetic materials. Given their artificial ferrite\nnature (Sec. III), ULMs can readily support this effect. The\nFaraday angle is given by [3], [8]\n\u0012F(z) =\u0000\u0012\f+\u0000\f\u0000\n2z\u0013\n;with\f\u0006=!p\n\u000f(\u0016\u0006\u0014);(6)\nwhere\u0016and\u0014are the Polder tensor components in (3b),\nwith the resonance ( !0) given by (5) and the saturation\nmagnetization frequency ( !m) discussed in Sec. III. Inter-\nesting, the ULM allows the option to reverse the direction\nof Faraday rotation by simple voltage control (instead of\nmagnet mechanical flipping in a conventional ferrite) using\nan antiparallel transistor pair load, as demonstrated in [42].\nFigure 5 shows a reflective Faraday ULM metasurface\n(ULMS) structure, based on the particle in Fig. 3, and re-\nsponse, initially reported in [32]. The results confirm that the\nULM works exactly as a ferrite, whose equivalent parameters\nare given in the caption.\nFig. 5. Reflective Faraday rotating ULMS based on the particle in Fig. 3 [32].\n(a) Perspective representation of the metasurface with rotated plane of polar-\nization. (b) Theoretical [Eq. (6)] and experimental polarization rotation angle\nversus frequency. Here, \u000fr= 2:6,!0=2\u0019= 7:42GHz (Bequiv.\n0= 0:265 T),\n!m=2\u0019= 28 MHz (\u00160Mequiv.\ns = 1 mT) and\u000b= 1:9\u000210\u00003Np\n(\u0001H=\u000b!0=\r\u0016 0= 0:4mT.)\nULM Faraday rotation has also been reported in transmis-\nsion, using the circular-slot ULM structure mentioned at the4\nend of Sec. III. Using slots, and hence equivalent magnetic\ncurrents, instead of rings supporting electric currents, that\nstructure really operates as an artificial magneto-optic material,\nwith a permittivity tensor replacing the magnetic tensor in (3a).\nA similar Faraday rotation effect may also be achieved using\narrays of twisted dipoles loaded by transistors [43].\nV. A PPLICATIONS\nA. Metasurface Isolators\nThe transmissive ULMS in [41] can be straightforwardly\napplied to build a Faraday isolator [3], [4], [44], [45], as shown\nin Fig. 6. As the wave propagates from the left to the right, its\npolarization is rotated 45\u000eby the left ULMS in the rotation\ndirection imposed by the transistors (here, clock-wise). It thus\nreaches the polarizer with its electric field perpendicular to\nthe conducting strips and therefore unimpededly crosses it. It\nis finally rotated back to its initial (vertical) direction by the\nright ULMS, whose rotation direction is opposite to the left\none (here, counter-clockwise). In the opposite direction, the\nright ULMS rotates the wave polarization in such a manner\nthat its electric field is parallel to the conducting strips of the\npolarizer, so that the wave is completely reflected. It is then\nrotated again by the right polarizer and gets back to the right\ninput orthogonal to the original wave7.\nFig. 6. Isolator using two transmissive Faraday rotation ULMSs [41] and a\n45\u000epolarizer. The bottom-left inset shows the transmissive “magneto-optic”\nslot-ULM demonstrated in [41], that may be used for this application.\nFaraday rotation is not the only approach to realize spatial\nisolation , as in Fig. 6. Such isolation may be simply achieved,\nwithout any gyrotropy but still magnetlessly, with a metasur-\nface consisting of back-to-back antenna arrays interconnected\nby transistors [47]; this nonreciprocal metasurface may exhibit\nan ultra wideband response and provide transmission gain.\nB. Nonreciprocal Antenna Systems\nThe ULM structure may be used in various nonreciprocal\nradiating (antenna, reflector and metasurface) applications.\nFigure 7 shows a nonreciprocal antenna system and its re-\nsponse [48]. The structure [Fig. 7(a)] is a ULM magnet-\nless version of the nonreciprocal ferrite-loaded Composite\nRight/Left-Handed (CRLH) open-waveguide leaky-wave an-\ntenna introduced in [49], [50], with the ferrite material re-\nplaced by a 1D ULM structure. This structure may be used\nas a nonreciprocal full-space scanning antenna [Figs. 7(b)\nand (c)], whose unidirectionality provides protection against\ninterfering signals, or as an antenna diplexer system , where\nnonreciprocity effectively plays the role of a circulator with\nhighly isolated uplink ( 3!1) and downlink ( 2!3) paths.\n7Lossy polarizers can be added, if necessary (The final wave could be\nreflected back), for dissipative (rather than reflective) isolation [46].\nFig. 7. ULM CRLH leaky-wave isolated-antenna or antenna-duplexer sys-\ntem [48]. (a) Prototype with port definitions. (b) Measured radiation patterns.\n(c) Measured scattering parameters.\nC. Isolators and Circulators\nULM technology also enables various kinds of nonre-\nciprocal components. Figure 8(a) shows a ULM microstrip\nisolator [38]. The ULM structure below the microstrip line is\ncomposed of two rows of transistor-loaded rings with opposite\nbiasing, and hence opposite allowed wave rotation directions.\nAs the wave from the microstrip line reaches a ring pair,\nits mode is coupled into a stripline mode with strip pair\nconstituted by the longitudinal sections of the overlapping\nrings, and usual antisymmetric currents. In the propagation\ndirection where these currents are co-directional, with allowed\nrotation direction of the ULM, the stripline mode is allowed\nto propagate, whereas in the opposite propagation direction, it\nis inhibited and dissipates in matching resistors on the rings.\nFigure 8(b) shows a ULM microstrip circulator [38], which is\nbased on mode-split counter-rotating modes as all circulators.\nFig. 8. ULM components [38]. (a) Isolator [51]. (b) Circulator [38].\nVI. C ONCLUSION\nWe have presented an overview of ULM structures and\napplications. The ULM physics has been described in great\ndetails, revealing that the ULM really represents an artificial\nferrite medium. It is in fact the only existing medium of\nthe kind. It has been pointed out that the ULM may offer\nunique extra benefits compared to ferrites, such a multiband\noperation, ultra broadband and electronic Faraday rotation\ndirection switching.5\nREFERENCES\n[1] A. G. Fox, S. E. Miller, and M. T. Weiss, “Behavior and applications\nof ferrites in the microwave region,” The Bell System Technical Journal ,\nvol. 34, no. 1, pp. 5–103, Jan 1955.\n[2] G. L. Matthaei, E. M. T. Jones, and S. B. Cohn, “A nonreciprocal,\nTEM-mode structure for wide-band gyrator and isolator applications,”\nIRE Transactions on Microwave Theory and Techniques , vol. 7, no. 4,\npp. 453–460, October 1959.\n[3] B. Lax and K. J. Button, Microwave ferrites and ferrimagnetics .\nMcGraw-Hill, 1962.\n[4] L. J. Aplet and J. W. Carson, “A Faraday effect optical isolator,” Applied\nOptics , vol. 3, no. 4, Apr. 1964.\n[5] S. Wang, J. Crow, and M. Shah, “Studies of magnetooptic effects\nfor thin-film optical-waveguide applications,” IEEE Transactions on\nMagnetics , vol. 7, no. 3, pp. 385–387, September 1971.\n[6] G. P. Rodrigue, “A generation of microwave ferrite devices,” Proc. IEEE ,\nvol. 76, no. 2, pp. 121–137, Feb. 1988.\n[7] T. Aoyama, T. Hibiya, and Y . Ohta, “A new Faraday rotator using a thick\nGd:YIG film grown by liquid-phase epitaxy and its applications to an\noptical isolator and optical switch,” Journal of Lightwave Technology ,\nvol. 1, no. 1, pp. 280–285, March 1983.\n[8] D. M. Pozar, Microwave Engineering , 4th ed. Wiley, 2011.\n[9] M. Faraday, Faraday’s Diary , T. Martin, Ed. George Bell and Sons,\n1933, vol. IV , Nov. 12, 1839 - June 26, 1847.\n[10] A. G. Gurevich and G. A. Melkov, Magnetization Oscillations and\nWaves . CRC Press, 1996.\n[11] L. Rayleigh, “On the magnetic rotation of light and the second law of\nthermo-dynamics,” Nature , vol. 64, pp. 577–578, Oct 1901.\n[12] L. D. Landau and L. P. Pitaevskii, Electrodynamics of Continuous Media ,\n2nd ed., ser. Course of theoretical physics. Butterworth-Heinemann,\n1984, vol. 8, Chap. XII.\n[13] A. K. Zvezdin and V . A. Kotov, Modern Magnetooptics and Magne-\ntooptical Materials . CRC Press, 1997.\n[14] C. Caloz, A. Al `u, S. Tretyakov, D. Sounas, K. Achouri, and Z.-L.\nDeck-L ´eger, “What is nonreciprocity? – Part I,” IEEE Antennas Wireless\nPropag. , vol. 17, 2018, this issue.\n[15] ——, “What is nonreciprocity? – Part II,” IEEE Antennas Wireless\nPropag. , vol. 17, 2018, this issue.\n[16] S. Tanaka, N. Shimomura, and K. Ohtake, “Active circulators amp, the\nrealization of circulators using transistors,” Proc. IEEE , vol. 53, no. 3,\npp. 260–267, March 1965.\n[17] Y . Ayasli, “Field effect transistor circulators,” IEEE Transactions on\nMagnetics , vol. 25, no. 5, pp. 3242–3247, Sep 1989.\n[18] I. J. Bahl, “The design of a 6-port active circulator,” in 1988., IEEE MTT-\nS International Microwave Symposium Digest , May 1988, pp. 1011–\n1014 vol.2.\n[19] S. Tang, C. Lin, S. Hung, K. Cheng, and Y . Wang, “Ultra-wideband\nquasi-circulator implemented by cascading distributed balun with phase\ncancelation technique,” IEEE Transactions on Microwave Theory and\nTechniques , vol. 64, no. 7, pp. 2104–2112, July 2016.\n[20] S. A. Ayati, D. Mandal, B. Bakkaloglu, and S. Kiaei, “Integrated\nquasi-circulator with RF leakage cancellation for full-duplex wireless\ntransceivers,” IEEE Transactions on Microwave Theory and Techniques ,\nvol. PP, no. 99, pp. 1–10, 2017.\n[21] K. Fang and J. F. Buckwalter, “A tunable 5–7 GHz distributed active\nquasi-circulator with 18-dBm output power in CMOS SOI,” IEEE\nMicrowave and Wireless Components Letters , vol. 27, no. 11, pp. 998–\n1000, Nov 2017.\n[22] S. Qin, Q. Xu, and Y . E. Wang, “Nonreciprocal components with\ndistributedly modulated capacitors,” IEEE Transactions on Microwave\nTheory and Techniques , vol. 62, no. 10, pp. 2260–2272, Oct 2014.\n[23] N. Estep, D. Sounas, and A. Al `u, “Magnetless microwave circulators\nbased on spatiotemporally modulated rings of coupled resonators,”\nTrans. Microw. Theory Tech. , vol. 64(2), pp. 502–518, 2016.\n[24] A. Kord, D. L. Sounas, and A. Al `u, “Magnet-less circulators based on\nspatiotemporal modulation of bandstop filters in a delta topology,” IEEE\nTransactions on Microwave Theory and Techniques , vol. 66, no. 2, pp.\n911–926, Feb 2018.\n[25] S. Taravati, N. Chamanara, and C. Caloz, “Nonreciprocal electromag-\nnetic scattering from a periodically space-time modulated slab and\napplication to a quasisonic isolator,” Phys. Rev. B , vol. 96, no. 16, pp.\n165 144:1–11, October 2017.\n[26] S. Taravati and C. Caloz, “Mixer-duplexer-antenna leaky-wave system\nbased on periodic space-time modulation,” IEEE Transactions on An-\ntennas and Propagation , vol. 65, no. 2, pp. 442–452, Feb 2017.[27] N. A. Estep, D. L. Sounas, J. Soric, and A. Al `u, “Magnetic-free non-\nreciprocity and isolation based on parametrically modulated coupled-\nresonator loops,” Nat. Phys , p. 923, Nov. 2014.\n[28] R. Negar and K. Harish, “Magnetic-free non-reciprocity based on\nstaggered commutation,” Nature Communications , vol. 7, p. 7:11217,\n2016.\n[29] A. L. Cullen, “A travelling-wave parametric amplifier,” Nature , vol. 181,\nno. 4605, p. 332, Feb. 1958.\n[30] R. Landauer, “Parametric amplification along nonlinear transmission\nlines,” J. Appl. Phys. , vol. 31, no. 3, pp. 479–484, Mar. 1960.\n[31] L. E. Franks and I. W. Sandberg, “An alternative approach to the\nrealization of network transfer functions: The N-path filter,” The Bell\nSystem Technical Journal , vol. 39, no. 5, pp. 1321–1350, Sept 1960.\n[32] T. Kodera, D. L. Sounas, and C. Caloz, “Artificial Faraday rotation using\na ring metamaterial structure without static magnetic field,” Appl. Phys.\nLett., vol. 99, no. 3, pp. 031 114:1–3, July 2011.\n[33] L. Landau and E. Lifshits, “On the theory of the dispersion of magnetic\npermeability in ferromagnetic bodies,” Physikalische Zeitschrift der\nSowjetunion. , vol. 8, pp. 153–169, 1935.\n[34] D. Sounas and A. Al `u, “Angular-momentum-biased nanorings to realize\nmagnetic-free integrated optical isolation,” ACS Photonics , vol. 1(3), pp.\n198–204, 2014.\n[35] D. L. Sounas, T. Kodera, and C. Caloz, “Electromagnetic modeling\nof a magnet-less non-reciprocal gyrotropic metasurface,” IEEE Trans.\nAntennas Propag. , vol. 61, no. 1, pp. 221–231, January 2013.\n[36] K. Achouri and C. Caloz, “Design, concepts and applications of elec-\ntromagnetic metasurfaces,” arXiv , vol. 1712.00618v1, Dec. 2017, in\npreparation.\n[37] D. Polder, “On the theory of ferromagnetic resonance,” Physica , vol. 15,\nno. 1, pp. 253–255, 1949.\n[38] T. Kodera, D. L. Sounas, and C. Caloz, “Magnetless non-reciprocal\nmetamaterial (MNM) technology: application to microwave compo-\nnents,” IEEE Trans. Microw. Theory Tech. , vol. 61, no. 3, pp. 1030–1042,\nMarch 2013.\n[39] L.-P. Carignan, V . Boucher, T. Kodera, C. Caloz, A. Yelon, and\nD. M ´enard, “Double ferromagnetic resonance in nanowire arrays,” Appl.\nPhys. Lett. , vol. 45, no. 6, pp. 062 504–1:3, Aug. 2009.\n[40] V . Boucher, L.-P. Carignan, T. Kodera, C. Caloz, A. Yelon, and\nD. M ´enard, “Effective permeability tensor and double resonance of\ninteracting bistable ferromagnetic nanowires,” Phys. Rev. B , vol. 80, pp.\n224 402:1–11, Dec. 2009.\n[41] T. Kodera, D. L. Sounas, and C. Caloz, “Faraday rotation by artificial\nelectric gyrotropy in a transparent slot-ring metamaterial structure,” in\nProceedings of the 2012 IEEE International Symposium on Antennas\nand Propagation , July 2012, pp. 1–2.\n[42] ——, “Switchable magnet-less non-reciprocal metamaterial (MNM) and\nits application to a switchable Faraday rotation metasurface,” IEEE\nAntenn. Wireless Propag. Lett. , vol. 11, pp. 1454–1457, December 2012.\n[43] Y . Shen, D. Kim, Ye, L. Wang, I. Celanovic, L. Ran, J. D. Joannopoulos,\nand M. Solja ˇci´c, “Metamaterial broadband angular selectivity,” Physical\nReview B , vol. 90, p. 125422, September 2014.\n[44] C. L. Hogan, “The ferromagnetic Faraday effect at microwave frequen-\ncies and its applications: The microwave gyrator,” The Bell System\nTechnical Journal , vol. 31, no. 1, pp. 1–31, Jan 1952.\n[45] R. Rosenberg, C. B. Rubinstein, and D. R. Herriott, “Resonant optical\nFaraday rotator,” Appl. Opt. , vol. 3, no. 9, pp. 1079–1083, Sep 1964.\n[46] A. Parsa, T. Kodera, and C. Caloz, “Ferrite based non-reciprocal radome,\ngeneralized scattering matrix analysis and experimental demonstration,”\nIEEE Trans. Antennas Propag. , vol. 59, no. 3, pp. 810–817, Mar. 2011.\n[47] S. Taravati, B. S. Khan, S. Gupta, K. Achouri, and C. Caloz, “Nonre-\nciprocal nongyrotropic magnetless metasurface,” IEEE Trans. Antennas\nPropag. , vol. 65, no. 7, pp. 3589–3597, June 2017.\n[48] T. Kodera, D. L. Sounas, and C. Caloz, “Non-reciprocal magnet-less\nCRLH leaky-wave antenna based on a ring metamaterial structure,”\nIEEE Antenn. Wireless Propag. Lett. , vol. 10, pp. 1551–1554, January\n2012.\n[49] T. Kodera and C. Caloz, “Uniform ferrite-loaded open waveguide\nstructure with CRLH response and its application to a novel backfire-to-\nendfire leaky-wave antenna,” IEEE Transactions on Microwave Theory\nand Techniques , vol. 57, no. 4, pp. 784–795, April 2009.\n[50] ——, “Low-profile leaky-wave electric monopole loop antenna using\nthe\f= 0 regime of a ferrite-loaded open waveguide,” IEEE Trans.\nAntennas Propag. , vol. 58, no. 10, pp. 3165–3174, Oct. 2010.\n[51] T. Kodera, D. L. Sounas, and C. Caloz, “Isolator utilizing artificial\nmagnetic gyrotropy,” in 2012 IEEE/MTT-S International Microwave\nSymposium Digest , June 2012, pp. 1–3." }, { "title": "2112.05961v1.Optical_excitation_of_electromagnons_in_hexaferrite.pdf", "content": " \n1 \n Optical excitation of electromagnon s in hexaferrite \nHiroki Ueda1,*, Hoyoung Jang2, Sae Hwan Chun2, Hyeong -Do Kim2, Minseok Kim2, Sang -\nYoun Park2, Simone Finizio1, Nazaret Ortiz Hernandez1, Vladimir Ovuka3, Matteo Savoini3, \nTsuyoshi Kimura4, Yoshikazu Tanaka5, Andrin Doll1, and Urs Staub1,* \n1Swiss Light Source, Paul Scherrer Institute, 5232 Villigen -PSI, Switzerland. \n2 PAL-XFEL, Pohang Accelerator Laboratory, Pohang, Gyeongbuk 37673, South Korea. \n3 Institute for Quantum Electronics, Physics Departmen t, ETH Zurich, 8093 Zurich, \nSwitzerland. \n4 Department of Advanced Materials Science, University of Tokyo, Kashiwa, Chiba 277 -8561, \nJapan. \n5 RIKEN SPring -8 Center, Sayo, Hyogo 679 -5148, Japan. \n \nAbstract : Understanding ultrafast magnetization dynamics on the microscopic level is of \nstrong current interest due to the potential for applications in information storage. In recent \nyears, the spin-lattice coupling has been recognized to be essential for ultrafast magnetization \ndynamics . Magnetoelectric multi ferroics of type II possess intrinsic correlation s among \nmagnetic sublattice s and electric polarization ( P) through spin -lattice coupling , enabling \nfundamentally coupled dynamics between spins and lattice . Here we report on ultrafast \nmagnetization dynamics in a room -temperature multiferroic hexaferrite possessing \nferrimagnetic and antiferromagnetic sublattices, revealed by time -resolved resonant x -ray \ndiffraction . A femtosecond above -bandgap excitation tri ggers a coherent magnon in which \nthe two magnetic sublattices entangle and give rise to a transient modulation of P. A novel \nmicroscopic mechanism for triggering the coherent magnon in this ferrimagnetic insulator \nbased on the spin-lattice coupling is prop osed. Our finding opens up a novel but general \npathway for ultrafast control of magnetism . \n* To whom correspondence should be addressed : hiroki.ueda@psi.ch and urs.staub@psi.ch \n \n2 \n Magnetoelectric multiferroics of type II commonly exhibit magnetic order that \ninduces ferroelectricity, and hence, there is naturally a strong coupling between magnetism \nand lattice, i.e., spin -lattice coupling . This class of materials has attracted enormous interest \nin the last two decades [ 1]. Interest in these multiferroics lies, for example, in electric \n(magnetic) manipulation of magnetism (polarization) [ 2-4] and in coupled spin-lattice \nexcitations called electromagnons [ 5,6]. This inherent correlation between magnetism and \nferroelectricity enables simultaneous control of principal excitation characters, magnons and \nphonons, in a wide dynamic range. However , time-resolved studies on the coupling, e.g., \noptical excitation of electro magnons and attempts to understand the microscopic origin of \ntheir creation, have been limited so far [7-10]. \nOptical manipulation of magnetic moments and identification of the underlying \nexcitation mechanism are important objectives towards new -generation storage technologies \nand have been intensely debated in the context of spin -lattice coupling . These are, e.g., \nultrafast non -thermal magnetization switching/reorientation via a change in \nmagnetocrystalline anisotropy [ 11-13], a coherent excitation of magnon s by driving an \nelectromagnon or optical phonon mode [ 8,14,15], and an ultrafast demagnetization that \ngenerates a transverse strain wave via the Einstein -de Haas effect [ 16]. The importance of \nspin-lattice coupling is well established in multiferroics with their non -trivial magnetic orders \nresult ing in remarkable magnetoelectric effects. Often , multiferroicity emerge s due to \nmagnetic frustration resulting in a high sensitivity of the property to bond angle s. Tuning a \nlattice parameter , e.g. , using strain , is indeed a standard approach to obtain multiferroicity \n[17]. The well -documented knowledge of spin -lattice coupling in multiferroics inspires us to \nexplore a new pathway to create magnons via ultrafast modulation in the lattice other than \nknown mechanisms, e.g., coherent displacive excitation [ 11] and the inverse Cotton -Mouton \neffect [ 18]. \nIn this paper, we study the ultrafast magnetization dynamic s in a room -temperature \nmultiferroic Z -type hexaferrite Sr 3Co2Fe24O41 with a ferrimagnetic (FM) and an \nantiferromagnetic (AFM) cycloidal sublattice [19,20]. The dynamics of the individual \nsublattices upon a femtosecond above -bandgap excitation were investigated by time -resolved \nresonant x -ray diffraction (tr -RXD). A coherent entangled magnon mode of the AFM and FM \nsublattices is observed and is described by the Landau -Lifshitz -Gilbert (LLG) equation as an \nelectromagnon mode . Nonreciprocal directional dichroism (NDD) in microwave transmission \nspectra further confirm s its origin as an electromagnon mode. An alternative excitation \nmechanism for the magnon creat ion based on the ultrafast Einstein -de Haas effect is \nproposed . Such a novel excitation mechanism for optical driving of magnetic modes possibly \nneeds to be more general ly considered in frustrated magnetic insulators . \nZ-type hexaferrite Sr 3Co2Fe24O41 crystals are commonly described in space group \nP63/mmc with lattice constants of a ≈ 5.87 Å and c ≈ 52.07 Å [ 19]. The crystal structure, \n3 \n displayed in Fig. 1a , comprises two basic magnetic blocks, S block s and L block s, which \nalternately stack along [001]. Because of many magnetic Fe sites in the chemical unit cell, the \nmagnetic structure of hexaferrites is described by a collinear -FM structure in each magnetic \nblock. Arrows in Figs. 1b and 1c represent the local structures: a red (blue) arrow represents \nthe magnetic moment of an L (S) block μL (μS). There is magnetic frustration at the interface \nbetween S and L blocks, which is tunable by chemical sub stitution that changes the specific \nFe-O-Fe bond angle (see the green d otted box in Fig. 1a ) [21]. As a result, a non -collinear \nmagnetic structure shown in Fig. 1b is stabilized below the multiferroic transition \ntemperature TMF ≈ 410 K [ 20,22], whereas a collinear -FM structure shown in Fig. 1c appears \nabove TMF. \n \nFigure 1 . Crystal structure and magnetic structure s of Z-type hexaferrite \nSr3Co2Fe24O41 and time -resolved resonant x -ray diffraction setup. (a) Crystal structure \nof the Z -type hexaferrite Sr 3Co2Fe24O41 drawn by VESTA [38]. Co2+ sits on some Fe3+ \nsites. A green dotted box shows an enlarged view of the local structure at the interface \nbetween adjacent magnetic blocks, where some atoms not relevant to the magnetic \nfrustration are omitted. ( b,c) Simplified magnetic structure s of Sr 3Co2Fe24O41 in (b) the \nconical phase and ( c) the collinear -ferrimagnetic phase. Red and blue arrows denote net \nmoments in magnetic L and S blocks ( μL and μS), respectively. In the black dotted box , the \nprecession of μL in the magnetic excitation mode is shown by black arrows . Here μS is \nomitted because of its small precession amplitude. ( d) Sketch of the experimental setup \nused for the time -resolved resonant diffraction experiments (details are described in \nMethod ). (e) Resonant diffraction profile fr om Sr 3Co2Fe24O41 along (00 L) in the conical \nphase. \n \nThe non -collinear transverse -conical structure is composed of a collinear -FM \ncomponent with a propagation vector k = (0, 0, 0) and a non -collinear cycloidal component \nwith k = (0, 0, 1). Hereafter, we refer to the former component as the FM compone nt and to \nthe latter as the AFM component. Whereas the FM component hosts magnetization ( M), the \n \n4 \n AFM component gives rise to electric polarization ( P) through spin -orbit couplings \n[19,23,24]. A previous RXD study on Sr3Co2Fe24O41 investigated the magnetic sublattices \nseparately using two reflections sensitive to the individual sublattices [24] [see Fig 1e for a \ntypical resonant diffraction profile along (00 L) in the conical phase] . It revealed a strong \ncoupling between the two sublattices in the adiabatic limit, essential for the magnetoele ctric \neffect. \nTHz time -domain spectroscopy (THz-TDS) revealed the presence of an \nelectromagnon mode around 1 THz [ 25,26], indicative of the magnetoelectric coupling on \nultrafast timescales. This mode is the higher -frequency mode of the two possible \nelectromagnon modes with AFM resonance obtained by solving the LLG equation. The \nlower -frequency mode though has not been reachable in the THz-TDS experiment. The \nexpected d ynamics of the slower mode are visualized in the d otted box of Fig. 1b , which \ninvolves anti -phase oscillation between adjacent two μL and also between two μS with two -\norder smaller motions for the latter . In addition, an even slower magnetic mode is observed \nby the microwave spectroscopy technique at GHz frequenc ies [27,28], which might be \nreferred to as the toroidal -magnon mode or Nambu -Goldstone mode [ 29], the lowest -energy \nmagnon mode in the transverse conical state . \n \nRESULTS \nWe collected intensities of the (003) space -group forbidden and of the (004) allowed \nBragg reflections sensitive to the AFM and the FM component, respectively, in addition to \nthe fluorescence signal. The optical pump beam wavelength was 400 nm, which resul ts in an \nabove -bandgap excitation. Figure 2 shows tr -RXD intensities of the (003) and (004) Bragg \nreflections measur ed with polarization at 709.6 eV , normalized by unperturbed intensities \n(IOFF). Following the laser excitation, both reflections show ( O-1) a rapid decrease within \n~200 fs (see Figs. 2 a and 2b), (O-2) oscillations (see Figs. 2 c and 2d), (O-3) a slow response \nthat is either a decrease [for (003) and (004) with low fluence] or an increase [for (004) with \nhigh fluence] (see Figs. 2 e and 2f), and ( O-4) a recovery following t he precedent transient \nchanges . The time traces of the normalized intensities (ION/IOFF) are fitted by these four \ncomponents defined as \n𝐼ON\n𝐼OFF(𝑡)=−𝐴rap[1−e−(𝑡−𝑡0)𝜏rap⁄]+𝐴osce−(𝑡−𝑡0)𝜏osc⁄cos[2𝜋𝑓𝑡−𝜑]−𝐴slow[1−\ne−(𝑡−𝑡0)𝜏slow⁄]+𝐴rec[1−e−(𝑡−𝑡0)𝜏rec⁄]+1. (1) \nHere Arap (rap), Aosc (osc), Aslow (slow), and Arec (rec) are amplitudes (relaxation times ) of ( O-\n1) the rapid decrease, ( O-2) oscillation , (O-3) slow response , and ( O-4) recovery , \nrespectively , f is the oscillation frequency of the mode , t0 is time zero, and is a phase shift. \nThe fit results in an oscillation frequency of f1 ≈ 42.0±1.0 GHz at room temperature ( shown \n5 \n as translucent blue curves in Fig. 2 c,d). Best f its for the parameters are listed in \nSupplementary Information. \nThe oscillations exhibit the same frequency for the two reflections, but they are \nopposite in phase. Because the (003) and (004) reflections sampl e the AFM and FM \ncomponents, respectively, the oscillations with the same frequency indicate a coherent \nexcitation of a magnon mode entangl ing both components. Cooling to 84 K shows \nqualitatively similar results except for; (I) f1 shifts to a slightly higher frequency [~46.3±0.3 \nGHz (see the black translucent curves in Figs. 2 c,d, representing the fitted oscillation \ncomponent)] , (II) the oscillation lives more protracted , and (III) the slow response of (004) \nturns from an increasing to decreasing behavior despite the equal fluence used (compare \nblack and da rk green curves ). \n \nFigure 2 . Time traces of (003) AFM and (004) FM reflections. (a,c,e) (003) AFM and \n(b,d,f) (004) FM reflections measured with polarization of incident x -ray beams at 709.6 \neV; (a,b) short time trace around t0 (from 0.5 ps to 1.5 ps ), (c,d) medium time range ( 5 \nps to 40 ps) to improve visibility of the intensity oscillations , and (e,f) extended time trace \n(from 50 ps to 200 ps) . Each color corresponds to a different laser fluence or temperature \ndepicted with the label of the same c olor, and solid curves are best fits. A translucent curve \nin (c) and ( d) shows the oscillation component in the best fit. All data were taken at room \ntemperature except for black curves taken at 84 K. Vertical and horizontal dotted lines \nindicate t0 and the intensities before t0, respectively. \n \nMagnetic c ircular dichroic signals of the (004) reflection can directly capture M [24] \nand its time evolution upon laser excitation. Figure 3 show s tr-RXD intensities taken with \ncircular polarization at room temperature and extracted circular dichroic signals . An \nadditional feature appears in the circular dichroic signals, ( O-5) a significantly slower \noscillation at f2 ≈ 5.7±0.1 GHz (see black data in Fig. 3a ). The observation of a rapid decrease \nwithin < 200 fs followed by an increase of the overall signals indicate s an ultrafast \ndemagnetization followed by a delayed and slower increase of the FM component. The \n \n6 \n increase in the FM component goes hand in hand with the reduction of the AFM component \n(see Fig. 2 e), suggesting a transformation of AFM to FM moment component s. \n \nFigure 3 . Time traces of the circular dichroic (004) FM reflection . (a) Extended time \ntrace with high fluence ( 50 ps to 500 ps, 5.3 mJ/cm2) and ( b) short time range around t0 \n(2 ps to 27 ps, 5.3 mJ/cm2 and 0.6 mJ/cm2), measured at room temperature . Red (blue) \nplots are normalized intensities taken with C+ (C), while black plots are the difference \nbetween them, i.e., IC+ IC. In ( b), data taken with C are vertically shifted by 6 [arb. \nunits] . Inset shows d ata taken with low fluence (0.6 mJ/cm2) for better visibility of the \noscillation with f1. Errors are shown as translucent filling in ( b). \n \nTo better capture the mode s observed in the time traces , we performed microwave \nspectroscopy measurements . Figure 4 shows the NDD signals of microwave transmission \nspectra in the relevant frequency range s taken at room temperature and in various magnetic \nfields up to ±0.35 T. Resonant feature s with clear NDD are noticeable in both frequency \nranges, consistent with f1 and f2. Furthermore, the NDD signals indicate both modes are \nelectric - and magnetic -dipole active [ 6], meaning that these modes have a n electromagnon \ncharacter . Note that the higher -frequency f1 mode exhibits extraordinar y large NDD of up to \n3.5 dB, which translates to a change in transmitted power > 100% and is much larger than the \nmaximum change of ~11% for the lower -frequency f2 mode . This implies a considerable \nresonator strength coupled to the electric field of the microwave and corresponding to \nsignificant modulation s in M and P when the mode is resonantly excited . While NDD is less \npronounced in the low -frequency range, microwave absorption is resonantly enhanced (see \nSupplementary Information ), reminiscent of previous NDD/ microwave studies on a chiral \nmagnet [30]. All modes harden for increasing magnetic fields; resonan ce frequenc ies in the \nlower -frequency regime are almost linear in the amplitude of the static magnetic field \nconsistent with the Zeeman effect, whereas that of the higher -frequency mode is strongly \nnon-linear. Furthermore, the modes soften for elevat ed temperatures, as shown in Fig. S6 (see \nSupplementary Information). These results indicate that the oscillation with f2 could be \nassigned to the toroidal magnon or Nambu -Goldstone mode , the lowest -energy magnon mode \n \n7 \n showing NDD [31], while the oscillation with f1 would be assigned to a more complicated \nelectromagnon excitation . Together with no evident heat -driven lattice expansion shown in \nSI, we conclude that the oscillations in tr -RXD are not caused by longitudinal strain waves \nbut are of magnetic origin. \n \nFigure 4. Nonreciprocal directional dichroic signals in microwave transmission \nspectra. Around (a) low - (3-20 GHz , f2) or (b) high - (35-45 GHz , f1) frequency range in \nvarious magnetic fields at room temperature with a coplanar waveguide setup (see \nMethods) . \n \nAs the slow non-oscillating response in tr-RXD intensities of the (004) reflection (O-\n3) qualitatively differs for lower and higher laser fluences, we varied the laser fluence on \n(003) and (004) reflections for a given time delay of 200 ps (see Fig. 5). On the one hand, the \nintensity of the (004) reflection decreases for increasing fluence up to a threshold laser \nfluence of approximately 0.7 mJ/cm2 at room temperature , above which the intensity starts to \nincrease and saturates at ~3.1 mJ/cm2 far above its initial value. In contrast, the (003) \nintensity decreases monotonically over the whole range. On the other hand , the tr -RXD \nintensities of both the reflections monotonically reduce for increasing laser fluence at 84 K \n(see blue data points i n Fig. 5). \nThis behavior can be explained by the absorbed energy of the optical excitation . When \nexcited at room temperature , the temperature increase s above TMF after equilibration . Above \nTMF, the AFM component is absent, but the FM component is enhanced . Note that the in -\nplane M in the multiferroic phase does not strongly depend on temperature [ 22], indicating \nthat any thermal effect on the FM component becomes relevant only for temperature s above \nTMF. This view is supported by the estimated lattice expansion shown in Fig. S1a , consistent \nwith a temperature of ~410 K [32], which is precisely TMF. Note also that there is a depth \nprofile of the deposited energy ( effective temperature ), and the estimated temperature is the \nmean value within the probe depth . Thus, the top layers have an effective temperature well \nbeyond TMF. This scenario explains both the suppression of the (003) reflection and the \nincrease of the (004) reflection. \n \n8 \n \nFigure 5. Fluence dependence of (003) AFM and (004) FM reflections . (a) (003) AFM \nand ( b) (004) FM measured , taken at room temperature (red) and 84 K (blue) , and at 200 \nps delay time . Dotted lines indicate normalization baselines. \n \nUnderstanding the fast dynamics occurring within the first ~200 fs after excitation , as \nshown in Figs. 2 a and 2b, is not straightforward as it is close to the time resolution of the \nbeamline [ 33]. Nevertheless, tr -RXD intensities for both AFM and FM components show a \nrapid decrease of magnetic signals within ~200 fs. Such a quick response cannot directly \nrelate to a lattice expansion but is possibly due to an electron ic change [ 34]. To gain more \ninformation, we collected time-resolved x-ray absorption spectra (tr-XAS) in fluorescence \nmode . Figure 6c shows tr -XAS signals for polarization below or above the prominent main \npeak of the Fe L3 edge 0.5 ps after the excitation (see Fig. 6a). At both photon energies, \ntransient signals show a rapid variation within ~200 fs, the same time scale as found in the \nresponse of the tr-RXD intensities. \nAs the fast response with a tiny amplitude is also present in the time-resolved circular \ndichroic signals of the (004) reflection sensitive to M (Fig. 3b ), the quick response is \nexpected to be relate d to the ultrafast demagnetization. Note that the rapid change in \nfluorescence signals results in an intermediate state with a long lifetime with a monotonic \namplitude increase for increasing laser fluence, as shown in Fig. 6b. This long -lived change \nin XAS remains up to 500 ps (see Fig. 6d), consistent with a magnetic origin , whereas we \n \n9 \n expect that an electronic change would be very short -lived [35]. It is likely caused by \nvariations in x-ray magnetic linear dichroism due to demagnetization (from the AFM and /or \nthe FM component) , though here the changes present only the variation of the single \npolarization channel (). The absence of a clear oscillation with the frequency of f2 implies \nthat this mode is unlikely the simple Kittel mode. However, a more detailed investigation \nwith improved statistic s would be required to clarify what exactly causes the tr -XAS signals. \n \nFigure 6. Time -resolved f luorescence signals. (a) Fluorescence spectra taken with normal \nincidence of -polariz ed x-ray beams for the excited (red, ON, 8.2 mJ/cm2, 0.5 ps after the \nexcitation) and unperturbed (blue, OFF) sample and the difference between them (black). \n(b) Fluence dependence o f normalized fluorescence signal s taken with the photon energy \nof 708.5 eV at a time delay of 0.5 ps. Time traces of normalized fluorescence signal s (c) \nwith a time range around t0 (1 ps to 1 ps) and ( d) up to 500 ps. Red, blue, and black \nsymbols were taken with the photon energy of 708.5 eV, 710.6 eV, and 708.9 eV, \nrespectively. The solid curves in ( c) and ( d) are represents the best fit to Eq. (1) excluding \nthe oscillation term. Note that the long -lived change indicates an additional recovery term \ntaking more than 500 ps. Horizontal and vertical dotted lines indicate a normalization \nbaseline and t0, respectively. \n \nDISCUSSION \nWe have found that there are; ( O-1) a rapid response within ~200 fs, observed in tr -\nRXD intensities of both the (003) and (004) reflections, a time-resolved circular dichroic \n \n10 \n contrast in (004) , as well as in the tr-XAS, ( O-2) oscillations with the frequency of f1 (≈ \n42.0±1.0 GHz at room te mperature) in tr -RXD intensities of both the reflections that are of \nmagnetic origin as supported by the microwave spectra , (O-3) a slow response subsequently \nfollowing the rapid one, which we interpret as energy transfer from the electronic to the \nlattice system, (O-4) a recovery that takes longer than 500 ps , and ( O-5) oscillations with the \nfrequency of f2 (≈5.7±0.1 GHz) in the time-resolved circular dichroic signals of the (004) \nreflection most likely representing the lowest -energy magnon mode. Here we discuss ( i) the \noscillation s with f1 represent ing a coherent magnon mode and ( ii) how an above -bandgap \nexcitation launches the magnon mode. \nI. Assignment of the entangled magnon mode \nWe assume a simple phenomenological model to describe the experimental ly \nobserved entangled magnon mode , i.e., the frequency of f1 with opposite phase oscillation s \nfor the two reflections . We consider two order parameters with temporal modulation s; an \nAFM order parameter 𝑆AFM(𝑡)=𝑆AFM[1+𝛿e𝑖(𝜔𝑡−𝛷)] and an FM order parameter 𝑆FM(𝑡)=\n𝑆FM[1−𝛿e𝑖(𝜔𝑡−𝜑)]. Here SAFM and SFM are static order parameters of the AFM component \nand the FM component, respectively, δ is the amplitude of temporal modulation, ω is an \nangular frequency, and Φ is an arbitrary phase. The (004) reflection at resonance involves \nscattering proces ses of magnetic, charge, or orbital (or electric quadrupole) origin. We refer \nto a parameter describing the latter two scatterings as SCO. \nThe form factor 𝐹̂ including the polarization dependence at resonance can be written \nas a tensor of the form \n𝐹̂=(𝐹𝜎−𝜎′𝐹𝜋−𝜎′\n𝐹𝜎−𝜋′𝐹𝜋−𝜋′). (2) \nAs the magnetic field in the tr -RXD experiments lies in both the scattering and the basal \nplane s, the magnetic part of the form factor Fm at (004), i.e., the FM component, is in the \nscattering plane [ 𝐅𝑚𝐐=∑𝐦𝑗exp(𝑖𝐐∙𝐫𝑗) 𝑗 , where mj and rj are the magnetic moment and \npositional vector of a magnetic atom j, respectively, and Q is the scattering vector.], whereas \nFm at (003), i.e., the AFM component, is normal to M. For polarized incoming x -rays, \nmagnetic scattering occurs in both the π -σ’ and π -π’ channels for the (003) reflection. As \nthere is no charge/orbital scattering contribution at (003), the (003) intensities are \nproportional to |𝐅𝑚(003)|2\n and thus |𝑆AFM|2 as \n𝐼(003)≈|𝑆AFM|2. (3) \nOn the contrary, the charge/orbital scattering at (004) appear s in the π -π’ chan nel for \npolarized incoming x -rays (see Supplementary Information ), whereas magnetic scattering \nappears in the π -σ’ channel because Fm is in the scattering plane . Due to the orthogonal \n11 \n relation between the π-π’channel ( charge/orbital scattering ) and π-σ’ channel ( the magnetic \nscattering ), the (004) intensities are \n𝐼(004)≈|𝑆FM|2+|𝑆CO|2=𝐼FM+𝐼CO. (4) \nTr-RXD intensities of the reflections are \n𝐼(003)(𝑡)≈|𝑆AFM(𝑡)|2=𝑆AFM∗𝑆AFM[1+𝛿e−𝑖(𝜔𝑡−𝜑)][1+𝛿e𝑖(𝜔𝑡−𝜑)] \n≈𝐼(003)[1+2𝛿cos(𝜔𝑡−𝜑)] (5) \nand \n𝐼(004)(𝑡)≈|𝑆FM(𝑡)|2+|𝑆CO|2=𝑆FM∗𝑆FM[1−𝛿e−𝑖(𝜔𝑡−𝜑)][1−𝛿e𝑖(𝜔𝑡−𝜑)]+|𝑆CO|2 \n≈𝐼FM[1−2𝛿cos(𝜔𝑡−𝜑)]+𝐼CO=𝐼(004)−2𝛿𝐼FMcos(𝜔𝑡−𝜑). (6) \nEquation s (5) and (6) describe an entangled mode of the two sublattices and reproduce the \nexperimental observation that is of a single frequency but opposite phase between the (003) \nreflection and (004) reflection. The initial phase of the oscillations shown in Fig. 2 result s in a \npositive sign of δ, representing an increase of the AFM component subsequent to the \nexcitation. Although chang es in magnetic exchange interactions through lattice \nthermalization can launch magnetic excitations, it should result in a negative sign of δ due to \nthe reduction of the AFM component at higher temperatures . \nUsing an effecti ve magnetic Hamiltonian and solving the LLG equation for the \nhexaferrite, Chun and coworkers obtained two k = 0 magnon modes with different \nfrequencies [ 26]. They found that the two magnetic moments in adjacent L blocks or S bl ocks \nhave antiphase motions in the magnon modes. The parameters that quantitatively reproduc e \nthe higher -frequency electromagnon mode observed by THz-TDS at low temperature (~1 \nTHz at 20 K), provide a slower electro magnon mode frequency of ~58 GHz. Th is frequency \nis close to f1 (≈46.3±0.3 GHz at 84 K) found in the tr -RXD intensities . This mode softens \nwhen increasing temperature s, for example, ~42.0±1.0 GHz at room temperature , resembling \nthe softening of the higher -frequency electromagnon mode by increa sing temperatures [26]. \nThis indicates that f1 obtained from the tr -RXD intensities represents the coherent excitation \nof the slower electro magnon mode with two entangle d sublattices as sketched in Fig. 1b that \nis domina ted by the μL motion. The antiphase motion of magnetic moments in adjacent L \nblocks or S blocks creates transient P along [001] through exchange striction, similar to the \nreported electromagnon mode [ 25,26]. Inducing transient P through exchange striction \nimplies a more intense resonator strength for the electromagnon mode than for the toroidal \nmagnon mode, which modulates static P originated from spin -orbit couplings , consistent with \nthe observed strengths in the NDD signals (see Fig. 4 ). \nII. Mechanism of magnon excitation \nThe known mechanisms for optical magnon excitations are based on a change in \neffective magnetic -field direction due to; ( i) a photomagnetic effect changing magnetic \ncrystalline anisotropy or known as a coherent displacive excitation [ 11] and (ii) the inverse \n12 \n Cotton -Mouton effect or known as an impulsive stimulated Raman scattering [ 18]. The \nantiphase spin motions require the opposite direction of transient effective magnetic field \nbetween adjacent two L blocks . Among the possible tensor components that can change the \neffective magnetic -field direction in su ch a way , however , the symmetry of the hexaferrite \ndoes not support such a transient change through these photomagnetic effects (see \nSupplementary Information for details). Furthermore, photothermalization changing magnetic \nexchange interactions is ruled o ut, as mentioned above. \nFollowing the intensive laser excitation, ultrafast demagnetization of a FM sublattice \noccurs as observed in a rapid drop of the transient (004) circular dichroic signal. It has been \nshown that m ost of the lost angular momentum can be transferred within 200 fs into the \nangular momentum of lattice , creating a transverse force that triggers a strain wave through \nthe ultrafast Einstein -de Haas effect [ 16]. In ferromagnets with uniform M, the force only \nappears at the surface because a microscopic torque for spatially uniform M cancels in bulk. \nOn the other hand, in ferrimagnets, possessing two blocks with opposite M as drawn in Fig. \n7a, the force has the same direction at the interface and is therefore maximal there. This will \ninitiate shear waves (see Fig. 7b and Supplementary Information ). \nThere i s the magnetic frustration at the interfaces of the two magnetic blocks, L and S, \nthat creates the non-collinear magnetic structure in the hexaferrite family. Chemical \nsubstitution and thus angle change s of specific iron -oxygen bonds (see Fig. 1a ) alter the \nmagnetic exchange interaction that strongly affect s the frustration and correspondingly the \nmagnetic structure . As the launched shear wave is on a timescale much faster (< 200 fs) than \nthe period of the magnetic excitation , it can launch the ele ctromagnon excitation owing to the \nsensitivity of the magnetic structure due to the tiny displacements of interfacial bonds . \nWe can estimate the amplitude of the spin precession from the oscillation amplitude \nof the tr -RXD intensities of the (003) AFM ref lection. The fitting results in Fig. 2 yield the \noscillation amplitude in normalized intensities as ~3% at room temperature with 0.5 mJ/cm2, \nand ~5% at 84 K and 1.2 mJ/cm2. This roughly corresponds to a precession angle of μL as \n~1.1° and ~1.8° at room temperature and 84 K, respectively (see Supplementary Information \nfor details). These angle s are smaller than that in the Y-type hexaferrite Ba0.5Sr1.5Zn2(Fe 1-\nxAlx)12O22 (x = 0.08) manifesting a different excitation mechanism, namely a coherent \ndisplacive excitation type, which is allowed by symmetry [ 10]. However , the amplitudes in \nangle displacement found in our experiment for the Z -type hexaferrite are comparable to the \nreported large -amplitude spin dynamics upon electromagnon resonance excitation in \nTbMnO 3 [8]. \n13 \n \nFigure 7. Sketch es of the temporal r esidual strain at the interface of adjacent \nmagnetic blocks due to angular momentum transfer from ultrafast demagnetization \nof the ferrimagnetic blocks ( the ultrafast Einstein -de Haas effect ). Ultrafast \ndemagnetization of the ferrimagnetic component , red a nd blue arrows in (a), resulting in \nangular momentum transfer , yellow and black arrows in ( a) and black circle arrows in (b), \nresulting in uncompensated residual strain at the interface of two magnetic blocks as \ndenoted by thick black arrows. \n \nIn summary, we investigat ed the ultrafast magneti zation dy namics of the room -\ntemperature multiferroic Z-type hexaferrite Sr 3Co2Fe24O41 that exhibit s two magnetic \nsublattices ( ferrimagnetic one and pure antiferromagnetic one ), utilizing time -resolved \nresonant x-ray diffraction upon a femtosecond above -bandgap excitation. Our observations \nshow a coherent entangled magnetic excitation of the two sublattices instantaneously \nfollowing the optical excitation , which reflects a simultaneous optical control of two \ncharacters in a multiferroic material, magnetization and electric polarization, i.e., an \nelectromagnon excitation. We propose that a direct tuning of magnetic frustration due to \nshear waves launched by ultrafast demagne tization through the Einstein -de Haas effect could \nbe the origin of the coherent excita tion of the electro magnon. This novel non-thermal \nmechanism can trigger transient spin modulation in materials with magnetic frustration. The \noptical manipulat ion of lattice, based on ultrafast demagnetization and shear waves via the \nEinstein -de Haas effect , could be, in general , a more competing channel to explore \ncorrelation s among electronic degrees of freedom and the lattice on an ultrafast timescale . \n \nMethods \nTime -resolved resonant x -ray diffraction. Tr-RXD experiments were performed at the \nRSXS end -station of the SSS beamline in the PAL -XFEL [ 33]. The photon energies of \nmonochromatic x -ray beams were in the vicinity of the Fe L3 edge (≈ 710 eV). The \npolarization of the beams was either linear (π) or circular (C+/C ) controlled by the insertion \nof a magnetized Fe-foil with in -plane M (±) at an angle of approximately 45º with respect to \n \n14 \n the x -ray beam, acting as a circular polarizer [ 36]. The circular polarization degree estimated \nfrom the transmitted intensities is ~80%. Switching M of the Fe -foil with an electromagnet \nreverses the circularly polarized component. The optical pump beams (400 nm) were p-\npolarization and collinear to the probe beam. A single -crystal of Sr 3Co2Fe24O41 was mounted \ntogether with a pair of permanent magnets (≈ 0.1 T along [100] and parallel to the scattering \nplane ) on the diffractometer implemented in the end -station (see Fig. 1d ). Two avalanche \nphotodiodes collected simultaneous diffraction and fluorescence signals. A gas -monitor \ndetector collected incident x -ray pulse intensities and normalized the signals on a shot-to-shot \nbasis. The penetration dept hs of the beams are estimated to be ~34 nm and ~13 nm for the \npump (obtained from ellipsometry measurements) and for the probe (obtained from x -ray \nabsorption spectroscopy measurements [ 37]), respectively. The measurements were \nperformed at either room temperature or base temperature of the LN 2-cooled cryostat ~84 K. \nThe repetition rates of x -ray probe and optical pump pulses were 60 Hz and 30 Hz , \nrespectively . The length of the x -ray pulses is ~80 fs and is used for convolution of each time \ntrace with the corresponding Gaussian function. \nMicrowave transmission spectroscopy. Room temperature microwave transmission \nexperiments were performed at the Paul Scherrer Institute and employed a coplanar \nwaveguide setup . The s ame crystal as used for the x-ray experiments was placed on top of the \ncoplanar waveguide, with the [001] direction along the microwave propagation direction. \nBefore the experiments, the sample was processed with magnetic/electric fields, so -called \nmagneto electric poling procedures, to create a single -domain multiferroic state as \nexemplified in Ref. [ 24]. An electromagnet provided a calibrated magnetic field up to 0.35 T \nalong the [100] direction, which realize d the Voigt geometry , typical for NDD studies with \ncoplanar waveguides [ 30,31]. In the frequency range below 20 GHz, a commercial vector \nnetwork analyzer recorded the transmitted microwave power . To reach frequencies up to 50 \nGHz, a custom microwave setup was used. The NDD of transmitted microwaves wa s \nobtained by taking the ratio of transmission signals with opposite magnetic field polarity. In \nparticular, data recorded for an up -sweep in the ma gnetic field was divided by data recorded \nfor the corresponding down -sweep. The Supplementary Information contains further \ntechnical details on the microwave setup , transmission data, and NDD at elevated \ntemperature s and a different magnetic field orientation. \n \nData availability \nExperimental and model data are accessible from the PSI Public Data Repository [ 39]. \n \nReferences \n15 \n 1. Fiebig, M., Lottermoser, T., Meier, D. & Trassin , M. The evolution of multiferroics. Nat. \nRev. 1, 16046 (2016). \n2. Kimura, T., Goto, T., Shintani, H., Ishizaka, K., Arima, T. & Tokura , Y. Magnetic control \nof ferroelectric polarization, Nature 426, 55-58 (2003) . \n3. Tokunaga, Y., Taguchi, Y., Arima, T. & Tokura, Y. Electric -field-induced generation and \nreversal of ferromagnetic moment in ferrites. Nat. Phys. 8, 838 -844 (2012) . \n4. Kocsis, V., Nakajima, T., Matsuda, M., Kikkawa, A., Kaneko, Y., Takeshima, J., \nKakurai, K., Arima, T., Kagawa , F., Tokunaga, Y., Tokura, Y. & Taguchi , Y. \nMagnetization -polarization cross -control near room temperature in hexaferrite single \ncrystal. Nat. Commun. 10, 1247 (2019) . \n5. Pimenov, A., Mukhin, A. A. , Ivanov, V. Yu. , Travkin, V. D. , Balbashov, A. M. & Loidl , \nA. Possible evidence for electromagnons in multiferroic manganites. Nat. Phys. 2, 97-100 \n(2006) . \n6. Takahashi , Y., Shimano , R., Kaneko , Y., Murakawa , H. & Tokura , Y. Magnetoelectric \nresonance with electromagnons in a perovskite helimagnet. Nat. Phys. 8, 121 -125 (2012) . \n7. Talbayev , D., Trugman , S. A. , Balatsky , A. V. , Kimura , T., Taylor , A. J. & Averitt , R. D. \nDetection of coherent magnons via ultrafast pump -probe reflectance spectroscopy in \nmultiferroic Ba 0.6Sr1.4Zn2Fe12O22. Phys. Rev. Lett. 101, 097603 (2008) . \n8. Kubacka , T., Johnson , J. A., Hoffmann , M. C. , Vicario , C., de Jong , S., Beaud , P., \nGrübel , S., Huang , S.-W., Huber , L., Pattey , L., Chuang , Y.-D., Turner , J. J., Dakovski , \nG. L. , Lee , W.-S., Schlotter , M. P. , Moore , R. G. , Hauri , C. P. , Koohpayeh , S. M. , \nScagnoli , V., Ingold , G., Johnson , S. L. & Staub , U. Large -amplitude spin dynamics \ndriven by a THz pulse in resonance with an electromagnon. Science 343, 1333 -1336 \n(2014) . \n9. Burn , D. M. , Zhang , S., Zhai , K., Chai , Y., Sun , Y., van der Laan , G. & Hesjedal , T. \nMode -resolved detection of magnetization dynamics using x -ray diffractive ferromagnetic \nresonance. Nano Lett. 20, 345 -352 (2020) . \n10. Jang, H., Ueda , H., Kim , H.-D., Kim , M., Shin , K. W. , Kim , K. H. , Park , S.-Y., Shin , H. \nJ., Borisov , P., Jang , D., Choi , H., Eom , I., Staub , U. & Chun , S. H. 4D visualization of \nthe photoexcited coherent magnon by an X -ray free electron laser. Preprint at \nhttp://arxiv.org/abs/2110.15626 . \n11. Hansteen , F., Kimel , A., Kirilyuk , A. & Rasing , T. Femtosecond photomagnetic \nswitching of spins in ferrimagnetic garnet films. Phys. Rev. Lett. 95, 047402 (2005) . \n12. Stupakiewicz , A., Szerenos , K., Afanasiev , D., Kirlyuk , A. & Kimel , A. V. Ultrafast \nnonthermal photo -magnetic recording in a transparent medium. Nature 542, 71-74 \n(2017) . \n13. Stupakiewicz , A., Davies , C. S. , Szerenos , K., Afanasiev , D., Rabinovich , K. S. , Boris , A. \nV., Caviglia , A., Kimel , A. V. & Kirilyuk , A. Ultrafast phononic switching of \nmagnetization. Nat. Phys. 17, 489 -492 (2021) . \n16 \n 14. Nova , T. F. , Cartella , A., Cantaluppi , A., Först , M., Bossini , D., Mikhaylovskiy , R. V. , \nKimel , A. V. , Merlin , R. & Cavalleri , A. An effective magnetic field from optically \ndriven phonons. Nat. Phys. 13, 132 -136 (2017) . \n15. Afanasiev , D., Hortensius , J. R., Ivanov , B. A. , Sasani , A., Bousquet , E., Blanter , Y. M. , \nMikhaylovskiy , R. V. , Kimel , A. V. & Caviglia , A. D. , Ultrafast control of magnetic \ninteractions via light -driven phonons. Nat. Mater. 20, 607 -611 (2021) . \n16. Dornes , C., Acremann , A., Savoini , M., Kubli , M., Neugebauer , M. J. , Huber , L., Lantz , \nG., Vaz , C. A. F. , Lemke , H., Bothschafter , E. M. , Porer , M., Esposito , V., Rettig , L., \nBuzzi , M., Alberca , A., Windsor , Y. W. , Beaud , P., Staub , U., Zhu , D., Song , S., \nGlownia , J. M. & Johnson , S. L. The ultrafast Einstein -de Haas effect. Nature 565, 209 -\n212 (2019) . \n17. Wadati , H., Okamoto , J., Garganourakis , M., Scagnoli , V., Staub , U., Yamasaki , Y., \nNakao , H., Murakami , Y., Mochizuki , M., Nakamura , M., Kawasaki , M. & Tokura , Y. \nOrigin of the large polarization in multiferroic YMnO 3 thin films revealed by soft - and \nhard-x-ray diffraction. Phys. Rev. Lett. 108, 047203 (2012). \n18. Kalashnikova , A. M. , Kimel , A. V. , Pisarev , R. V. , Gridnev , V. N. , Usachev , P. A. , \nKirilyuk , A. & Rasing , Th. Impulsive excitation of coherent magnons and phonons by \nsubpicosecond laser pulses in the weak ferromagnet FeBO 3. Phys. Rev. B 78, 104301 \n(2008) . \n19. Kitagawa , Y., Hiraoka , Y., Honda , T., Ishikura , T., Nakamura , H. & Kimura , T. Low-\nfield magnetoelectric effect at room temperature. Nat. Mater. 9, 797 -802 (2010) . \n20. Soda , M., Ishikura , T., Nakamura , H., Wakabayashi , Y. & Kimura , T. Magnetic ordering \nin relation to the room -temperature magnetoeletric effect of Sr 3Co2Fe24O41. Phys. Rev. \nLett. 106, 087201 (2011) . \n21. Kimura , T. Magnetoelectric hexaferrites, Annu. Rev. Condens. Matter Phys. 3, 93-110 \n(2012) . \n22. Nakajima , H., Kawase , H., Kurushima , K., Kotani , A., Kimura , T. & Mori , S. \nObservation of magnetic domain and bubble structures in magnetoelectric Sr 3Co2Fe24O41. \nPhys. Rev. B 96, 024431 (2017) . \n23. Chai , Y. S. , Chun , S. H. , Cong , J. Z. & Kim, K. H. Magnetoelectricity in multiferroic \nhexaferrites as understood by crystal symmetry analyses. Phys. Rev. B 98, 104416 (2018) . \n24. Ueda , H., Tanaka , Y., Wakabayashi , Y. & Kimura , T. Insights into magnetoelectric \ncoupling mechanism of the room -temperature multiferroic Sr 3Co2Fe24O41 from domain \nobservation. Phys. Rev. B 100, 094444 (2019) . \n25. Kadlec , F., Kadlec , C., Vít, J., Borodavka , F., Kempa , M., Prokleška , J., Buršík , J., \nUhrecký , R., Rols , S., Chai , Y. S. , Zhai , K., Sun , Y., Drahokoupil , J., Goian , V. & \nKamba , S. Electromagnon in the Z -type hexaferrite (Ba xSr1-x)3Co2Fe24O41. Phys. Rev. B \n94, 024419 (2016) . \n17 \n 26. Chun , S. H. , Shin , K. W. , Kim , H. J., Jung , S., Park , J., Bahk , Y.-M., Park , H.-R., \nKyoung , J., Choi , D.-H., Kim , D.-S., Park , G.-S., Mitchell , J. F. & Kim, K. H. \nElectromagnon with sensitive terahertz magnetochroism in a room -temperature \nmagnetoelectric hexaferrite. Phys. Rev. Lett. 120, 027202 (2018) . \n27. Ebnabbasi , K., Mohebbi , M. & Vittoria , C. Magnetoelectric effects at microwave \nfrequencies on Z -type hexaferrite. Appl. Phys. Lett. 101, 062406 (2012) . \n28. Laguta , V., Kempa , M., Bovtun , V., Buršík , J., Zhai , K., Sun , Y. & Kamba , S. \nMagnetoelectric coupling in multiferroic Z -type hexaferrite revealed by electric -\nfieldmodulated magnetic resonance studies. J. Mater. Sci. 55, 7624 -7633 (2020) . \n29. Miyahara , S. & Furukawa , N. Theory of magneto -optical effects in helical multiferroic \nmaterials via toroidal magnon excitation. Phys. Rev. B 89,195145 (2014) . \n30. Okamura , Y., Kagawa , F., Mochizuki , M., Kubota , M., Seki , S., Ishiwata , S., Kawasaki , \nM., Onose , Y. & Tokura , Y. Microwave magnetoelectric effect via skyrmion resonance \nmodes in a helimagnetic multiferroic. Nat. Commun. 4, 2391 (2013). \n31. Iguchi , Y., Nii, Y. & Onose , Y. Magnetoelectrical control of nonreciprocal microwave \nresponse in a multiferroic helimagnet. Nat. Commun. 8, 15252 (2017) . \n32. Takada , Y., Nakagawa , T., Fukuta , Y., Tokunaga , M., Yamamoto , T. A. , Tachibana , T., \nKawano , S., Igawa , N. & Ishii, Y. Temperature dependence of magnetic moment \norientation in Co 2Z-type hexaferrite estimated by high -temperature neutron diffraction. \nJpn. J. Appl. Phys. 44, 3151 -3156 (2005) . \n33. Jang, H., Kim , H.-D., Kim , M., Park , S. H. , Kwon , S., Lee , J. Y., Park , S.-Y., Park , G., \nKim, S., Hyun , H.-J., Hwang , S., Lee , C.-S., Lim , C.-Y., Gang , W., Kim , M., Heo , S., \nKim, J., Jung , G., Kim , S., Park , J., Kim , J., Shin , H., Park , J., Koo , T.-Y., Shin , H.-J., \nHeo, H., Kim , C., Min , C.-K., Han , J.-H., Kang , H.-S., Lee , H.-S., Kim , K. S. , Eom , I. & \nRah, S. Time -resolved resonant elastic soft x -ray scattering at Pohang Accelerator \nLaboratory X -ray Free Electron Laser. Rev. Sci. Instrum. 91, 083904 (2020) . \n34. Bigot , J.-Y., Vomir , M. & Beaurepaire , E. Coherent ultrafast magnetism induced by \nfemtosecond laser pulses. Nat. Phys. 5, 515 -520 (2009) . \n35. Uemura , Y., Ismail , A. S. M., Park , S. H. , Kwon , S., Kim , M., Niwa , Y., Wadati , H., \nElnaggar , H., Frati , F., Haarman , T., Höppel, N., Huse , N., Hirata , Y., Zhang , Y., \nYamagami , K., Yamamoto , S., Matsuda , I., Katayama , T., Togashi , T., Owada , S., \nYabashi , M., Halisdemir , U., Koster , G., Yokoyama , T., Weckhuysen , B. M. & de Groot , \nF. M. F. Femtosecond charge density modulation in photoexcited CuWO 4. J. Phys. Chem. \nC 125, 7329 -7336 (2021). \n36. Kortright , J. B., Kim , S.-K., Warwick , T. & Smith , N. V. Soft x -ray circular polarizer \nusing magnetic circular dichroism at the Fe L 3 line. Appl. Phys. Lett. 71, 1446 -1448 \n(1997) . \n18 \n 37. Ueda , H., Tanaka , Y., Wakabayashi , Y., Tsurumi , J., Takeya , J. & Kimura , T. Multiple \nmagnetic order parameters coexisting in multiferroic hexaferrites resolved by soft x rays. \nJ. Appl. Phys. 128, 174101 (2020) . \n38. Momma , K. & Izumi , F. VESTA 3 for three -dimensional visualization of crystal, \nvolumetric and morphology data. J. Appl. Crystallogr. 44, 1272 -1276 (2011) . \n39. DOI: https://doi.org/10.16907/ efb03af6 -f28b-414f-8eb1 -77b31a035fb7 . \n \nAcknowledgements \nWe thank M. Burian for his advice in data analysis. The time -resolved resonant x -ray \ndiffraction experiments were performed at the RSXS end -station of the SSS beamline in the \nPAL -XFEL under proposal No. 2020 -2nd-SSS-011. The static resonant x -ray diffraction \nexperiments were performed at the X11MA beamline in the Swiss Light Source during in -\nhouse access and at the BL17SU in the SPring -8 under proposal No. 20180021. H.U. was \nsupported by the National Centers of Competence in Research in Molecular Ultrafast Science \nand Technology (NCCR MUST -No. 51NF40 -183615 ) from the Swiss National Science \nFoundation and from the European Union’s Horizon 2020 research and innovation \nprogramme under the Marie Skłodowska -Curie Grant Agreement No. 801459 – FP-\nRESOMUS. H.J. acknowledges the support by the National Research Found ation grant \nfunded by the Korea government (MSIT) (Grant No. 2019R1F1A1060295). S.H.C. was \nsupported by National Research Foundation of Korea (2019R1C1C1010034 and \n2019K1A3A7A09033399). N. O. H. was supported by the Swiss National Science \nFoundation ( No. 2 00021_169017 ). A.D. acknowledges funding from the PSI -internal CROSS \ninitiative. T.K. and Y.T. were supported by JSPS KAKENHI Grant Number JP19H00661. \n \nAuthor contributions \nH.U. and U.S. conceived and designed the project . H.U. and T.K. prepared a single \ncrystal of Sr 3Co2Fe24O41. H.U., N.O.H., Y.T., and U.S. performed synchrotron measurements \nto characterize the sample. S.F. fabricated the circular polarizer. H.U., H.J., S.H.C., H. -D.K., \nand U.S. performed the time -resolved resonant x -ray diffraction experiment, and H.U. \nanalyzed the experimental data. S. -Y.P. established the data acquisition system for the \nbeamline experiment. M.K. prepared the optical laser pump setup. V.O. and M.S. measured \noptical ellipsometry of the sample. A.D. built up the microwave spectroscopy setup and \nperformed the experiment. H.U., A.D., and U.S. interpreted the experimental results and \nwrote the manuscript. All authors contributed to its final version. \n 1 \n Supplementary Information for \nOptical excitation of electromagnons in hexaferrite \nHiroki Ueda1,*, Hoyoung Jang2, Sae Hwan Chun2, Hyeong -Do Kim2, Minseok Kim2, Sang -\nYoun Park2, Simone Finizio1, Nazaret Ortiz Hernandez1, Vladimir Ovuka3, Matteo Savoini3, \nTsuyoshi Kimura4, Yoshikazu Tanaka5, Andrin Doll1, and Urs Staub1,* \n1Swiss Light Source, Paul Scherrer Institute, 5232 Villigen -PSI, Switzerland. \n2 PAL-XFEL, Pohang Accelerator Laboratory, Pohang, Gyeongbuk 37673, South Korea. \n3 Institute for Quantum Electronics, Physics Department, ETH Zurich, 8093 Zurich, \nSwitzerland. \n4 Department of Advanced Materials Science, University of Tokyo, Kashiwa, Chiba 277 -8561, \nJapan. \n5 RIKEN SPring -8 Center, Sayo, Hyogo 679 -5148, Japan. \n \n* To who m correspondence should be addressed: hiroki.ueda@psi.ch and urs.staub@psi.ch \n 2 \n Contributions of orbital scattering \nIn general, scattering from an aspherical electron density created by the occupation \nvalence electrons in the orbitals (orbital scattering) can contribute to diffraction intensities at \nresonance. Here, w e perform a symmetry analysis to clarify if orbital scattering contributes to \nthe (003) and (004) reflec tions. Orbital scattering is described by the anisotropic x -ray \nsusceptibility tensor 𝑓̂. The tensor elements are restricted by the local (site) symmetry of the \nresonant atom, and the sum of the respective tensors with their phase factor gives a form \nfactor and intensity [ 1]. \nWe denote the symmetric tensor 𝑓̂as \n𝑓̂=(𝑓𝑥𝑥𝑓𝑥𝑦𝑓𝑥𝑧\n𝑓𝑥𝑦𝑓𝑦𝑦𝑓𝑦𝑧\n𝑓𝑥𝑧𝑓𝑦𝑧𝑓𝑧𝑧) (S1) \nin a Cartes ian coordinate system where x is along [100], y is along [120], and z is along [001]. \nThere are 10 Fe sites in a Z -type hexaferrite [ 2], and their local symmetries are summarized \nin Table S1. All Fe sites except the Fe4 and Fe8 sites have the three -fold rotational symmetry \nC3 along [001]. The local symmetry requires the relation 𝑓̂=𝐶3𝑓̂𝐶3−1, which results in fxx = \nfyy and fxy = fyz = fzx = 0. The symmetry -adopted 𝑓̂ is then \n𝑓̂=(𝑓𝑥𝑥00\n0𝑓𝑥𝑥0\n00𝑓𝑧𝑧). (S2) \nThe Fe4 and Fe8 sites contain the mirror symmetry m that is along <100> . The symmetry -\nadopted tensor follows the relation 𝑓̂=𝑚𝑓̂𝑚−1 and is represented as \n𝑓̂=(𝑓𝑥𝑥00\n0𝑓𝑦𝑦𝑓𝑦𝑧\n0𝑓𝑦𝑧𝑓𝑧𝑧). (S3) \nFor the sites lacking local C3 symmetry, i.e., the Fe (4) and Fe (8), there are three atoms at the \nsame z coordinate due to the global C3 symmetry. For (00L) reflections, only the z coordinate \nis relevant , allowing to define a n average 𝑓̂ for these three atoms that is equivalent to Eq. S2. \nAs the residual 𝑓̂’s are isotropic in the basal plane, the form factor 𝐹̂ for the (003) \nreflection is zero while that for the (004) reflection is \n𝐹̂=(𝐹𝑥𝑥00\n0𝐹𝑥𝑥0\n00𝐹𝑧𝑧). (S4) \nEquation S4 implies that orbital scattering appears only in the polarization -unrotated \nchannels, i.e., σ -σ’ and π -π’ and is independent of the azimuthal angle. \n 3 \n Table S1. Ten Fe sites in a Z -type hexaferrite with their Wyckoff positions and local \nsymmetries, referring to Ref. [2]. \nFe site Wyckoff position Local symmetry \nFe(1) 2a 3̅m. \nFe(2) 4f 3m. \nFe(3) 4e 3m. \nFe(4) 12k .m. \nFe(5) 4e 3m. \nFe(6) 4f 3m. \nFe(7) 4f 3m. \nFe(8) 12k .m. \nFe(9) 4f 3m. \nFe(10) 2c 6̅m2 \n \n 4 \n Lattice expansion due to photothermalization \nTo test if the Bragg peak position (2 θ) shifts due to heat -driven lattice expansion we \ncollected (00 L) diffraction profiles in the vicinity of the (004) reflection at several delay times \nfollowing laser excitation . Figure S1 shows peak shifts extracte d from Gaussian fits of the \n(004) reflection profiles reflecting possible changes in the lattice constant c. There is a tiny \nchange around ~4 ps after the laser excitation, and recovery sets in on a timescale of 500 ps \nsubsequently following the expansion. These peak shifts are more than an order of magnitude \nsmaller than the peak width and therefore do not result in observable variation in the intensity \nof the peak maxima used in the traces. The peak width of the (004) reflection gets broader by \nthe laser excitation again at ~4 ps after the laser excitation (see Fig. S1 ). However, these \nvariations are clearly too small to affect the observed oscillations in tr -RXD. \n \nFigure S1. Fitted results of 2 θ/θ profiles around (004) measure d. (a) Peak center in 2 θ \nand ( b) full width at half maximum (FWHM) as a function of time delay. Red and blue \nplots indicate data for the excited (ON) and unperturbed (OFF) samples, re spectively, at \nroom temperature, whereas black plots indicate a change in the lattice parameter c by the \nexcitation. The inset shows 2 θ/θ profiles at 40 ps of time delay. Laser fluence is 5.7 \nmJ/cm2, and dotted lines indicate either a baseline or t0. \n \n5 \n Comparison with other known mechan isms of optically excited coherent magnons \nHere we discuss the observed coherent magnon with the other known mechanism s, \ni.e., (i) a photomagnetic effect changing magnetic crystalline anisotropy or known as a \ncoherent displacive excitation [ 3] and (ii) the inverse Cotton -Mouton effect or known as an \nimpulsive stimulated Raman scattering [ 4]. These mechanisms are based on a change in \neffective magnetic -field direction since fo r an effective magnetic field Heff along the initial \nmagnetization M, the Landau -Lifshitz equation is \nd𝐌\nd𝑡=−𝛾𝐌×𝐇eff=𝟎, (S5) \nwhere γ is the gyroscopic ratio. Therefore, we consider if the mechanisms can generate Heff \nthat changes the magnetic -field direction. Besides, t he magnon mode involves two -order \nlarger motions for μL than μS, and so we more specifically consider if the mechanisms can \ngenerate Heff driving μL. To launch the antiphase motion of magnetic moments, Heff needs to \npoint in opposite directions between adjacent L blocks. \n(i) The photo -induced anisotropy field δ Hi through a coherent displacive excitation is \nphenomenologically described as \nδ𝐇𝑖(0)=𝜒̂𝑖𝑗𝑘𝑙𝐄𝑗(𝜔)𝐄𝑘∗(𝜔)𝐌𝑙(0), (S6) \nwhere 𝜒̂𝑖𝑗𝑘𝑙 is a fourth -rank polar tensor that is symmetric for j and k, E is the electric field of \nlight [ 3]. We use the Cartesian coordinate where x // [100] (= 1), y // [120] (= 2), and z // \n[001] (= 3). In our experimental setup, the scattering plane is spanned by the z-x plane, μL is \nin the x-y plane, and the pump beam is p-polarized. With this condition, p ossible components \nthat explain the antiphase spin motion are twelve components am ong 𝜒̂𝑖𝑗𝑘𝑙 (χ2111, χ2131, χ2331, \nχ3111, χ3131, χ3331, χ1112, χ1132, χ1332, χ3112, χ3132, and χ3332). \nΤhe symmetry of the L block belongs to point group 6̅m2 and allows only χ3131 among \nthe possible components. However, the symmetry postulates the component invariant \nbetween two adjacent L blocks, connected by the two -fold symmetry along [100]. Therefore, \nthe component can give only an in -phase but not an antiphase spin motion between two \nadjacent L blocks. \n(ii) The inverse Cotton -Mouton effect in a magnetic media is due to a photo -induced \nmodulation in dielectric permittivity that is proportional to M and is phenomenologically \ndescribed as \n𝐇𝑖ICME(0)=𝑔̂𝑖𝑗𝑘𝑙𝐄𝑗(𝜔)𝐄𝑘∗(𝜔)𝐌𝑙(0). (S7) \nHere 𝑔̂𝑖𝑗𝑘𝑙 is a fourth -rank polar tensor that is symmetric for j and k and represents a photo -\ninduced modulation in the permittivity of the media, and HiICME is the magnetic field induced \nby the Cotton -Mouton effect [ 5]. As 𝑔̂𝑖𝑗𝑘𝑙 has the same intrinsic symmetry with 𝜒̂𝑖𝑗𝑘𝑙, the \ninverse Cotton -Mouton effect can also not explain the observed antiphase spin motion. 6 \n Estimation of strain at the interface of magnetic blocks \n(i) Local symmet ry of magnetic blocks and elastic stiffness tensor \nAt first, we discuss the local symmetry of the magnetic blocks to find the elastic \nstiffness tensor components that are allowed by symmetry. The crystal structure of a Z -type \nhexaferrite comprises two magnetic blocks, an S block and an L block, as shown in Fig. 1(a) . \nAn S block locate s at z = 0 or 1/2 and is alternating with an L block located at z = 1/4 or 3/4. \nThe space group of the crystal structure is P63/mmc , and the point group is 6/ mmm . The point \ngroup of an S block is 3̅m, while that of an L block is 6̅m2. Note that these point groups are \nsubgroups of 6/ mmm . The symmetry -adapted elastic stiffness tensor Cij of the point groups \nare \n𝐶𝑖𝑗=\n( 𝐶11\n𝐶12\n𝐶13\n𝐶14\n0\n0 𝐶12\n𝐶11\n𝐶13\n−𝐶14\n0\n0 𝐶13\n𝐶13\n𝐶33\n0\n0\n0 𝐶14\n−𝐶14\n0\n𝐶44\n0\n0 0\n0\n0\n0\n𝐶44\n𝐶14 0\n0\n0\n0\n𝐶14\n𝐶112⁄−𝐶122⁄) \n (S8) \nfor 3̅m and \n𝐶𝑖𝑗=\n( 𝐶11\n𝐶12\n𝐶13\n0\n0\n0 𝐶12\n𝐶11\n𝐶13\n0\n0\n0 𝐶13\n𝐶13\n𝐶33\n0\n0\n0 0\n0\n0\n𝐶44\n0\n0 0\n0\n0\n0\n𝐶44\n0 0\n0\n0\n0\n0\n𝐶112⁄−𝐶122⁄) \n (S9) \nfor 6̅m2 [6]. \n \n(ii) Strain through the ultrafast Einstein -de Haas effect \nFollowing Ref. [ 7], we estimate the strain at the interface of two magnetic blocks. The \nultrafast demagnetization causes a transient volume torque density τ described as \n𝛕=−1\n𝛾d𝐌\nd𝑡 (S10) \n[8]. Based on a continuum model, τ contributes to the off -diagonal components of a \nmagnetization -dependent antisymmetric stress tensor 𝜎𝑀̂ as 𝜎12𝑀=−𝜎21𝑀=𝜏3, 𝜎23𝑀=−𝜎32𝑀=\n𝜏1, and 𝜎31𝑀=−𝜎13𝑀=𝜏2. The structural dynamics follows the equation of motion with a \ngiven stress tensor 𝜎̂ \n𝜌𝜕2𝑢𝑖\n𝜕𝑡2=∑𝜕𝜎𝑖𝑗\n𝜕𝑥𝑗𝑗 , (S11) \nwhere ρ is the mass density of the material, xj (j = 1, 2, 3) spans the Cartesian coordinate, and \nui is the displacement along xi. Here 𝜎̂ contains three contributions written as 7 \n 𝜎𝑖𝑗=∑𝐶𝑖𝑗𝑘𝑙𝜂𝑘𝑙 𝑘𝑙 +𝜎𝑖𝑗𝑀+𝜎𝑖𝑗Ext., (S12) \nwhere 𝜎𝑖𝑗Ext. is an external stress tensor component, and 𝜂𝑘𝑙 is the strain described as \n𝜂𝑘𝑙=1\n2(𝜕𝑢𝑘\n𝜕𝑥𝑙+𝜕𝑢𝑙\n𝜕𝑥𝑘), (S13) \nwhere k, l = 1, 2, 3. \nWe describe the tr -RXD setup as x1 // M and x3 // [001], namely τ = (τ 1, 0, 0). Then \nonly 𝜎23𝑀=−𝜎32𝑀=𝜏1 are non -zero in 𝜎𝑀̂. Here we ignore the diagonal components of 𝜎𝑀̂ \nsince such components do not contain a torque density. From Eqs. ( S10) and ( S11), we find, \nfor example, \n𝜌𝜕2𝑢2\n𝜕𝑡2=𝜕𝜎23\n𝜕𝑥3=𝜕𝜏1\n𝜕𝑥3=−1\n𝛾𝜕\n𝜕𝑥3(d𝑀1\nd𝑡), (S14) \nindicating that the spatial derivative of the ultrafast demagnetization accelerates th e \ndisplacement. The displacements are allowed only along x2 // [120] when we assume the \npump area is significantly larger than the probe area. To simpl ify the discussion, we assume \nthat the magnetization and ultrafast demagnetization are both uniform in a magnetic block. \nWithin t his assumption , the strain in the interior of a magnetic block is uniform. Besides, the \nforces from the internal torques cancel ev erywhere except at the interfaces of the two \nmagnetic blocks, similar to the case of the uniformly magnetized Fe film [ 7]. The stress \nthrough the torque density in a magnetic block appears at both the interfaces with adjacent \nmagnetic blocks, with the same amplitude but opposite sign. Assuming an interface of two \nmagnetic blocks is identical with the surface of the material, we have the same boundary \ncondition at the interfaces with the surface; \n𝜎3𝑗=0. (S15) \nThis means at the interfaces \n∑𝐶32𝑘𝑙𝜂𝑘𝑙 𝑘𝑙 =−𝜎32𝑀=−1\n𝛾d𝑀1\nd𝑡. (S16) \nThe elastic stiffness tensor is C32kl = C4j (j = k if k = l, otherwise j = 9kl). From Eqs. ( S9) \nand ( S16), we find only C44 = C3232 is allowed in the left -hand side of Eq. ( S16) for both S \nand L blocks. Therefore, the transverse strain at the two interfaces that each magnetic block \nhas is \n𝜂32=−1\n2𝛾𝐶3232d𝑀1\nd𝑡. (S17) \nSince the circular dichroic signals on the (004) FM reflection are proportional to \nmagnetization [ 9], the ultrafast demagnetization is quantitatively estimated from the rapid \ndrop of the signals displayed in Fig. S2 . The lost magnetization within 200 fs upon the optical \nexcitation with fluence of 5.3 mJ/cm2 is ~8% of the initial value, corresponding to 0.96 \nμB/f.u. as the magnetization at 1 kG is ~12 μB/f.u. (see Fig. S3 ) We s uppos e that the timescale \nand the ratio of angular momentum transfer to the lattice through the ultra fast Einstein -de 8 \n Haas effect are the same as those of the Fe film, shown in Ref. [ 7]; 80% of the lost angular \nmomentum transfers to the lattice within 200 fs. There are two interfaces between the \nmagnetic blocks for each formula unit. Thus, each interface acquires 0.38 μB/f.u. within 200 \nfs. Note that our discussion is based on the simpl e model with uniform magnetization in each \nmagnetic block , meaning a local ferromagnetic structure but not a ferrimagnetic structure as \nessentially the hexaferrite is. \nUsing the reported values of γ (= 2.76 π106 /s/T) and C3232 (= 3.761010 J/m3) of a \nsimilar hexaferrite [ 10,11], we obtain the strain η32 ≈ 3.510-2 and the estimated transverse \ndisplacements at the interfaces are ~2.2 pm , which is the comparable scale with observed \nlongitudinal phonon modes in, e.g., a Bi or Fe film [12, 7]. \nFig. S2. Time traces of the (004) FM reflection with circular polarization of incident x -rays \naround t0 measured at room temperature . The fluence is 5.3 mJ/cm2. Red and blue plots are \nnormalized intensities taken with C+ and C while black plots are the difference between \nthem. Data taken with C are vertically shifted to match its baseline with data taken with C+ . \n \nFig. S3. Magnetiza tion curve measured at room temperature. An applied magnetic field is \nparallel to the basal plane. \n9 \n Estimation of spin -precession angle in the electromagnon \nWe here quantify the spin -precession angle in the electromagnon mode upon the \nfemtosecond laser excitation. Based on the model shown in Fig. 1(b) , the transient magnetic \nform factor for (003) Fm(003)(t) is \n𝐅𝑚(003)(𝑡)=2(0\n−𝑖𝜇L(cosΔsin𝛽+sinΔcos𝛽cos𝜔𝑡)\n𝜇Ssin𝛼−𝑖𝜇LsinΔsin𝜔𝑡), (S18) \nwhere μL(S) and β(α) are the amplitude of the magnetic moment in an L (S) block and the half \nopening angle of the conical structure of an L (S) block, respectively (see Fig S4). Δ stands \nfor the angle between the transient magnetic moment of an L block and its equilibrium \ndirection (see Fig. 1b ), and ω is the angular frequency of the electromagnon mode. Here we \nuse the Cartesian coordinate where x is parallel to M, z is parallel to the scattering vector, and \ny is perpendicular to both x and z (see Fig. 1d for the experimental setup) . The wave vectors \nof incoming x -ray beam ( q) and outgoing x -ray beam ( q’) are then \n𝐪=(cos𝜃\n0\n−sin𝜃),𝐪′=(cos𝜃\n0\nsin𝜃), and 𝐪′×𝐪=(0\nsin2𝜃\n0), (S19) \nwith θ being the Bragg angle of the (003) reflection. For incoming π x-ray polarization, \nmagnetic scattering occur s in the πσ’ ( Iπσ’) and the ππ’ channels ( Iππ’), which are described by \n𝐼𝜋𝜎′=|𝐅𝑚(003)∙𝐪|2\n and 𝐼𝜋𝜋′=|𝐅𝑚(003)∙(𝐪′×𝐪)|2\n, (S20) \nrespectively. Using Eqs. ( S18)-(S20), the tr -RXD intensities of (003) are obtained as \n𝐼(003)(𝑡)≈4sin2𝜃(4𝜇L2cos2𝜃sin2𝛽+𝜇S2sin2𝛼)+2𝜇L2sin22𝜃sin2Δsin2𝛽cos𝜔𝑡,\n (S21) \nwhich is qualitatively the same as Eq. 5. Thus, the normalized intensity ION/IOFF is \n𝐼ON(𝑡)\n𝐼OFF=1+2𝜇L2cos2𝜃sin2Δsin2𝛽\n4𝜇L2cos2𝜃sin2𝛽+𝜇S2sin2𝛼cos𝜔𝑡, (S22) \nand Δ is directly defined by the oscillation amplitude extracted from the fits. \nFig. S4. Definition of the angles in the magnetic structure ; (a) α [(b) β] is the half opening \nangle of an S [L] block cone. \n \n10 \n Fitting of tr -RXD intensities \nTable S2 shows refined parameters using Eq. 1 convoluted with the time resolution to \ndescribe the tr-RXD intensities of the (003) reflection and (004) reflection following the \noptical excitation with fluence of 0.5 mJ/cm2 at room temperature. The d escription of each \nparameter is given in the main text. \n \nTable S2. Refined parameters f rom the tr -RXD intensities following the optical excitation \nwith fluence of 0.5 mJ/cm2 at room temperature. Strong correlation s among parameters \nprevent obtaining unique parameters for ( O-3) and ( O-4), which are omitted here. \nReflection Arap Aosc τrap [ps] τosc [ps] f [GHz ] φ [] \n(003) 1.4±0.010-2 3.0±0.010-2 0.25±0.04 18±2 42.0±1.0 179±4 \n(004) 8±110-2 4±110-2 0.13±0.02 32±6 40.8±0.9 1±1 \n \n 11 \n Setup for microwave transmission measurements \nThis section provides details on the experimental setup used for the microwave \ntransmission measurements. As stated in the main text, the sample was mounted on top of a \ncoplanar waveguide (see also Fig. S8 below) . For optimum perf ormance up to high \nmicrowave frequencies, a commercial coplanar waveguide board was used (Model B4350 -\n30C-50, Southwest Microwave) with suitable 2.92 mm coaxial connectors (Model 1093 -01A-\n5, Southwest Microwave) . Experiments in the sub -20 GHz regime utilized a commercial \nvector network analyze r (Model 8720ES, Agilent Technologies) with output power set to 1 \nmW. This vector network analyzer (VNA) directly yielded the microwave transmission \nparameter S21 as a function of microwave frequency f. In the sub -20 GHz regime, primary \nexperimental data were therefore transmission parameters S21(f) from 1 to 20 GHz recorded at \ndifferent magnetic field strengths. The magnetic field was stepped in 10 mT steps along a full \ncycle over its maximum range of ±0.35 T. Since the spectral features in recorded data were \nquite broad, a smoothing procedure with 240 MHz span was implemented in post -processing \nto reduce noise contributions . \nTo reach frequencies up to 50 GHz, a custom setup that implements microwave \nfrequency multiplication was built. In particular, the output frequency of a microwave source \n(Model EraSynth+, Era Instruments ; max. 15 GHz ) was multiplied by four. Two frequency \ndoublers (Model TB -973-CY244C +, Mini -Circuits) realized the four -fold frequency \nmultiplication. After passing the first frequency doubler, sufficient drive level and signal \npurity for the second frequency doubler w ere provided by a driving amplifier (Model EVAL -\nHMC383LC4 , Analog Devices ) followed by a filter (Model TB-883-1832C+ , Mini -Circuits) . \nMicrowave power levels were measured with a power detector (Model ZV47 -K44RMS+ , \nMini -Circuits). To keep the high -frequency signal path as short as possible, the second \nfrequency doubler and the power detector were directly connected to the coplanar waveguide \nprobe. To ease the voltage readout of the power detector, a modulation scheme with a lock -in \namplifier (Model elockin 203, Anaftec Instruments) was implemented: The microwave \nexcitation power was modulated sinus oidally at 910 Hz and the power detector output was \ndemodulated accordingly. The filter time -constant upon demodulation was set to 0.1 s and a \nsettling time of 0.6 s preceded any power level readout. Primary experimental data of this \nhigh frequency setup a re therefore the transmitted power Ptr at a particular frequency and a \nparticular magnetic field. At each microwave frequency, the same magnetic field cycle as for \nthe sub -20 GHz experiments was used. \nThe high -frequency setup had a lower -frequency cutoff around 30 GHz due to the \namplification and filtering stage in -between the two frequency doublers. The upper frequency \ncutoff was around 51 GHz due to high -frequency limitations of the utilized components. In \norder to verify the operation of the frequency m ultiplication chain around 40 GHz, we have \ntemporarily added an extra bandpass filter in front of the coplanar waveguide probe (Model 12 \n FB-3700 , Marki Microwave) . This filter exclusively passed the frequency range from 33 – 41 \nGHz and approved the features observed in Fig. 4b within the filter passband. In order to \naccess intermediate frequencies between 20 and 30 GHz, the setup was operated with only \none of the two frequency doublers. \n 13 \n \nMicrowave t ransmission data \nIn this section, transmission data related to the NDD data in the main text are shown. \nAn important aspect for transmission data is the normalization, which allows compensating \nfor instru mental features . For sub -20 GHz data, normalization was achieved by e ither data at \nthe highest field for frequencies below 16.7 GHz or data at zero field elsewhere. The \nparticular choice of the frequency crossover at 16.7 GHz was due to the resonance crossing \nthe instrumental baseline. Figure S5a shows the normalized transmission data obtained in this \nway. Corresponding NDD data is shown in Fig. 4a in the main text. Except for clearly visible \nfringes in the transmission data, the absorption peaks show a similar field-frequ ency trend as \nthe NDD data. By comparing the data at positive and negative fields, one can also infer the \nstronger attenuation for negative fields due to NDD. \nTo facilitate peak extraction, an additional normalization was applied to the signal \nattenuation |S21|n – 1: For each frequency f > 5.75 GHz, the attenuation was renormalized to \nthe mean peak attenuation for f > 5.75 GHz , where peak attenuation refers to the maximum \nattenuation over the scanned magnetic field cycle . In this way, frequency -dependent fringes \nwere reduced to a constant attenuation value, as is evident in Fig. S5 b. \nNote that contrary to transmission data, there is no need for any special normalization \nfor NDD extraction, since we extracted NDD by changing the direction of the magnetic field. \nWe have obtained equivalent results by extracting NDD upon changing the microwave \npropagation direction. However, ch anging the magnetic field direction yielded superior data \nquality, since the underlying data originated from exactly the same microwave transmission \npathway. \nFig. S5. Normalized transmission data |S21|n (f) at different magnetic fields . (a) Data \nnormalized to instrumental background extracted at highest field ( f < 16.7 GHz) and zero -\nfield (f > 16.7 GHz). (b) Additional scaling of attenuation for fringe suppression in order to \nease peak extraction. \n14 \n In the higher frequency range around 40 GHz, data normalization was less evident \nsince there was no clear absorption peak as in the sub-20 GHz data. Primary un -normalized \ndata are shown in Fig. S6 a, where the voltage reading of th e power detector was scaled by the \nnominal detector characteristics, i.e. a 34.4 dB/V slope and an output voltage of 1.0 V for a -5 \ndBm input signal. For frequencies around 39 GHz, a pronounced asymmetry along the field \naxis is already visible in this prim ary dataset. Normalized data are shown in Fig. S6b , where \nthe normalization is the maximum power level over field at each frequency. For negative \nfields, the heat map resembles ordinary microwave transmission data with a visually \ntraceable field -frequency evolution of the absorption peak. For positive fields, however, the \nNDD -induced asymmetry renders the absorption peak difficult to trace. In this case, resonant \nNDD is thus more pronounced than resonant absorp tion. This is quite different to the sub -20 \nGHz data, where resonant absorption is predominant and resonant NDD is only visible as a \nsmall additional change ( Fig. S6a). Note that for the sub -20 GHz data in Fig. S5a , a \nmaximum -based normalization as applied here around 40 GHz yields a qualitatively \ncomparable heat map . \nIn analogy to the sub -20 GHz data, we have verified that we obtain the same NDD \ncharacteristics by reversing the microwave propagation direction. With the setup at 40 GHz, \nthis implied to flip the direction of the coplanar waveguide probe. \nFig. S6. Transmission power Ptr (f) around 40 GHz at different magnetic fields . (a) Primary \npower levels. (b) Data normalized to maximum power level over field at each frequency. \n \n \n15 \n Microwave data at elevated temperature \nIn order to support the softening of the 40 GHz mode with temperature observed in \ntime-resolved data, additional microwave transmission spectra at elevated temperature were \nrecorded. For this purpose, a heat gun set to 80 °C was fixed in proximity of the s ample. A \nthermometer brought in contact with the sample displayed a temperature of 76 °C, which we \nconsider as an upper limit for the actual sample temperature . \nIn the sub -20 GHz regime, softening with temperature was approved by following the \nevolution o f the absorption peak in microwave transmission ( Figs. S7a) at room (blue) and \nelevated temperature s (orange) . It is readily seen that for a given magnetic field, heating \nresult ed in mode softening. Moreover, the effect progresse d with magnetic field strength. The \ncorresponding NDD spectra at room and elevated temperature s are shown in Figs. S7b and \nS7c, respectively. For comparison, the evolution in the absorption peak of the related \ntransmission spectra is superimposed o nto the NDD spectra as a black curve . Note that \nsudden jumps in the absorption peak evolution are due to the apparent dual -mode nature of \nthe recorded absorption and NDD spectra. Especially at the largest fields, the branch with \nslightly higher resonance frequency bec ame dominant (see Fig. S5 ). \nIn the 40 GHz regime, temperature effects were traced by following the field -\nfrequency evolution of the most intense NDD (Fig. S7 d) at room (blue) and elevated \ntemperature s (orange). Also in this case, the experimental data indicate field -progressive \nmode softening upon heating , even though less pronounced as in sub -20 GHz data . The \ncorresponding NDD spectra at room and elevated temperature are shown in Figs. S7e and \nS7f, respectively. When compared to the NDD spectrum in Fig 4b in the main text, NDD was \nweaker in these experiments. For further reference, the evolution of the NDD peak of the \ndataset in the main text is shown in green in Fig. S7 . We tentatively attribute these \ndifferences in NDD to the fact that the sample was re -mounted onto the coplanar wav eguide \nprobe in -between these two experiments. We believe that the geometry, orientation and \ndegree of electric polarization of the sample face that is exposed to microwaves are critical \nparameters for the resultant microwave properties. Importantly, these observations warrant \nfurther microwave studies of Sr3Co2Fe24O41 with a prospect of even larger NDD effects. \nLikewise, variations in NDD strength were observed among different runs with re -positioned \nsamples in the sub -20 GHz regime . However, NDD strengths stayed within ±0.3 dB with \nrespect to the data showed in Figs. 4 and S7. For sub -20 GHz data acquired at a different \nmagnetic field orientation, however, quite important changes in NDD were observed upon r e-\npositioning (see next section). 16 \n \nFig S7. Change in microwave resonances upon heating in the sub -20 GHz (a -c) and 40 GHz \n(d-f) frequency range. (a) Field -frequency evolution of absorption peak in transmission data \nat room temperature (blue) and elevated temperature (orange). There are two curves at each \ntemperature , field up - and down -sweep , and peak extraction was according to Fig. S5b. (b,c) \nNDD data at room and elevated temperature s, respectively. The superimposed black curve s \nare the absorption peaks shown in panel (a). Data in panel (b) is the same as in Fig. 4a in the \nmain text. (d) Field -frequency evolution of resonant NDD peak at room (blue) and elevated \ntemperature s (orange). The green curve depicts the NDD peak for the NDD spectrum shown \nin Fig. 4b in the main text. (e,f) NDD data at room and elevated temperature s, respectively. \n \n \n \n17 \n Microwave data at different field orientation \nIn previous sections, u ltrafast time -resolved X FEL and frequency -domain microwave \ntransmission experiments were performed with the static magnetic field along the [100] \ndirection. Since the two techniques bear different excitation mechanisms, we also performed \nmicrowave experiments with the field along the [120] direction. The orientation of the sa mple \nwith respect to the field and coplanar waveguide is depicted in Figs. S8a and S8b. The \nphotograph in Fig. S8c shows the actual setup with field along the [100] direction and the \nmicrowave propagation direction kμw as indicated. Rotation of the magnetic field around \n[001] was accomplished by fixation of the coplanar waveguide probe at a different angle. \nSuch a rotation does not only change the orientation of the relevant magnetic field in the \nbasal plane, but also changes the relative orientation between static and microwave fields. \nFigure S8d illustrates schematically the direction of the electric and magnetic microwave \nfields due to the coplana r waveguide. Changes observed upon sample rotation are thus a \ncombination of (i) anisotropy with respect to the external field within the basal plane and (ii) \nanisotropy with respect to the microwave excitation fields. \n \nFig S8. Sample and probe geometry for microwave transmission e xperiments. (a, b) \nSchematic layout of sample (black) placed on top of coplanar transmission waveguide cross \nsection (beige) with the external field along the [100] or [120] direction, respectively. (c) \nPhotograph of the sample m ounted on top of the coplanar waveguide. The center trace has a \nwidth of 43 mils (1.1 mm). (d) Indicative illustration of typical microwave field distribution \nwithin the waveguide cross section. The arrows represent field lines of constant intensity. \n \nFigure S9 shows microwave tr ansmission (a,c) and NDD (b,d) in the sub -20 GHz \nregime with field along the [120] direction. Two different datasets are shown in order to point \nat the subtle influence of sample positioning and microwave properties. For upper dataset, \nremarkably large NDD was observed for frequencies around 17.5 GHz . For the lower dataset, \nthe NDD pattern is in qualitative agreement, but at much lower intensity. In accordance to \nour observations in the 40 GHz regime ( Fig. S7), we anticipate the possib ility to fine tune the \nNDD response and reach even larger effects with Sr3Co2Fe24O41. In contrast to the 40 GHz \n18 \n data, however, sub-20 GHz data feature pronounced resonant microwave absorption that can \nbe compared against each other via the corresponding transmission spectra in panels (a) and \n(c). Interestingly, these transmission data show quantitatively comparable absorption leve ls. \nThe apparent differences in the color scale are mainly caused by the different NDD response s \nthat superimpose on normalized transmission data. \nAs is evident when comparing Fig. S9 and Fig. S5, there are important changes when \nflipping the magnetic field direction. To further ease the comparison, the field -frequency \nevolution of the principal absorption peak from Fig. S5b was superimposed as black line onto \ntransmission spectra in Fig. S9. These superimposed lines follow the highest -frequency ridge \nin the transmission spectra. We thus tentatively conclude that there is a common mode that is \nexcited with the field along either [100] or [120]. With the field along [100], this common \nmode is the lowest -frequency ridge in the transmission spectrum, while with the field along \n[120], the common mode becomes the highest -frequency ridge in the transmission spectrum. \nIt is also interesting to note that the with the field along [1 20], the lowe st-frequency ridge has \na resonance frequency of 7 GHz at 0.1 T, which is very close to the beating observed in laser -\nexcited time-domain XMCD data in the main text . \nFor further insight into the origin of the observed anisotropy in microwave \ntransmission, additional experiments with the magnetic field at intermediate orientations \nbetween [100] and [120] have been performed (data not shown). These experiment s indicated \nthat field orientations along [100] or [120] constitute extremum constellatio ns, from which \nwe tentatively deduce a two -fold rotational symmet ry of the apparent anisotropy. As \nmentioned above, the anisotropy can originate from both anisotropy related to the sample or \nrelated to the microwave excitation geometry . Since the sample ha s hexagonal symmetry in \nthe basal plane, we consider anisotropy with respect to the microwave excitation fields to be \nthe more likely cause for the two -fold rotational symmetry. Within this viewpoint , there is the \npossibility that laser -excited microwave mode in time -domain expe riments with the field \nalong [100] corresponds to the lowest -frequency ridge of the microwave -excited modes with \nthe field along [120] . Given the anisotropy and multitude of microwave -excited modes in the \nsub-20 GHz regime, however, the microscopic relationship between the microwave - and \nlaser -excited modes is beyond the scope of the current study and subject to further research. \nAn anisotropy in microwave properties was also observed for the mode around 40 \nGHz. In this case, the NDD response with field along the [100] direction ( Fig. S10 a) changed \nits characteristic field -frequency evolution when applying the field along the [120] direction \n(Fig. S10 b). As is re adily observed, the resonance was almost field-independent in this \nconfiguration. Note that the broad er range scans in Fig. S10 were performed with a ten times \nlower resolution in the microwave frequency axis than in other illustrated NDD spectra at 40 \nGHz. 19 \n \nFig S9. Microwave response in sub -20 GHz regime with magnetic field along [120], showing \nnormalized microwave transmission (a,c) and NDD (b,d) of the same sample mounted twice . \nFig S10. 40 GHz NDD with 1 GHz frequency steps and magnetic field along [100] (a) or \n[120] (b). \n \n20 \n References \n1. Dmitrienko , V. E. Forbidden reflections due to anisotropic x -ray susceptibility of crystals. \nActa Cryst. A39, 29-35 (1983) . \n2. Tachibana , T., Nakagawa , T., Takada , Y., Izumi , K., Yamamoto , T. A. , Shimada , T. & \nKawano , S. X-ray and neutron diffraction studies on iron -substituted Z -type hexagonal \nbarium ferrite: Ba 3Co2-xFe24+xO41 (x = 0-0.6). J. Magn. Magn. Mater. 262, 248 -257 \n(2003) . \n3. Hansteen , F., Kimel , A., Kirilyuk , A. & Rasing , T. Femtosecond photomagnetic \nswitching of spins in ferrimagnetic garnet films. Phys. Rev. Lett. 95, 047402 (2005) . \n4. Kalashnikova , A. M. , Kimel , A. V. , Pisarev , R. V. , Gridnev , V. N. , Usachev , P. A. , \nKirilyuk , A. & Rasing , Th. Impulsive excitation of coherent magnons and phonons by \nsubpicosecond laser pulses in the weak ferromagnet FeBO 3. Phys. Rev . B 78, 104301 \n(2008) . \n5. Shen , L. Q. , Zhou , L. F. , Shi, J. Y., Tang , M., Zheng , Z., Wu , D., Zhou , S. M. , Chen , L. \nY. & Zhao , H. B. Dominant role of inverse Cotton -Mouton effect in ultrafast stimulation \nof magnetization precession in undoped yttrium iron garnet films by 400 -nm laser pulses. \nPhys. Rev. B 97, 224430 (2018) . \n6. Gallego , S. V. , Etxebaria , J., Elcoro , L., Tasci , E. S. & Perez -Mato , J. M. Automatic \ncalculation on symmetry -adapted tensors in magnetic and non -magnetic materials: new \ntool of the Bilbao Crystallographic Server. Acta Cryst. A75, 438 -447 (2019) . \n7. Dornes , C., Acremann , A., Savoini , M., Kubli , M., Neugebauer , M. J. , Huber , L., Lantz , \nG., Vaz , C. A. F. , Lemke , H., Bothschafter , E. M. , Porer , M., Esposito , V., Rettig , L., \nBuzzi , M., Alberca , A., Windsor , Y. W. , Beaud , P., Staub , U., Zhu , D., Song , S., \nGlownia , J. M. & Johnson , S. L. The ultrafast Einstein -de Haas effect. Nature 565, 209 -\n212 (2019) . \n8. Einstein , A. & de Haas , W. J. Experimenteller Nachweis der Ampèreschen \nMolekularströme. Verhandl. Deut. Phys. Ges. 17, 152 -170 (1915). \n9. Ueda , H., Tanaka , Y., Wakabayashi , Y. & Kimura , T. Insights into magnetoelectric \ncoupling mechanism of the room -temperature multiferroic Sr 3Co2Fe24O41 from domain \nobservation. Phys. Rev. B 100, 094444 (2019) . \n10. Gonzalez , A. V., Lütgemeier , H. & Zinn , W. Field dependence of the NMR frequency in \nBaFe 12O19. Sol. Stat. Commun. 17, 1599 -1601 (1975) . \n11. Gudkov , V. V. , Sarychev , M. N. , Zherlitsyn , S., Zhevstovskikh , I. V., Averkiev , N. S. , \nVinnik , D. A. , Gudkova , S. A. , Niewa , R., Dressel , M., Alyabyeva , L. N. , Gorshunov , B. \nP. & Bersuker , I. B. Sub-lattice of Jahn -Teller centers in hexaferrite crystal. Sci. Rep. 10, \n7076 (2020) . \n12. Fritz , D. M. , Reis , D. A. , Adams , B., Akre , R. A. , Arthur , J., Blome , C., Bucksbaum , P. \nH., Cavalieri , A. L. , Engemann , S., Fahy , S., Falcone , R. W. , Fuoss , P. H. , Gaffney , K. J., \nGeorge , M. J. , Hajdu , J., Hertlein , M. P. , Hillyard , P. B. , Horn -von Hoegen , M., 21 \n Kammler , M., Kaspar , J., Kienberger , R., Krejcik , P., Lee, S. H., Lindenberg , A. M. , \nMcFarland , B., Meyer , D., Montagne , T., Murray , E. D. , Nelson , A. J., Nicoul , M., Pahl , \nR., Rudati , J., Schlarb , H., Siddons , D. P. , Sokolowski -Tinten , K., Tschentscher , Th., von \nder Linde , D. & Hastings , J. B. Ultrafast bond softening in bismuth: mapping a solid’s \ninteratomic potential with x -rays. Science 315, 633 -636 (2007) . \n \n " }, { "title": "1711.10790v1.Thermal_contribution_to_the_spin_orbit_torque_in_metallic_ferrimagnetic_systems.pdf", "content": "1 \n Thermal contribution to the spin -orbit torque in metallic/ferrimagnetic systems \nThai Ha Pham1, S.-G. Je1,2, P. Vallobra1, T. Fache1, D. Lacour1, G. Malinowski1, M. C. Cyrille3, G. Gaudin2, O. \nBoulle2, M. Hehn1, J.-C. Rojas -Sánchez1* and S. Mangin1 \n1 . Institut Jean Lamour, CNRS UMR 7198, Université de Lorraine, F-54011 Nancy, France \n2 . CNRS, SPINTEC, F -38000 Grenoble \n3. Leti, technology research institute, CEA, F-38000 Grenoble \n*juan -carlos.rojas -sanchez@univ -lorraine.fr \n \nAbstract \nWe report a systematic study of current -induced perpendicular magnetization switching in \nW/Co xTb1-x/Al thin films with strong perpendicular magnetic anisotropy . Various Co xTb1-x \nferrimagnetic alloys with different magnetic compensation temperatures are presented. The \nsystem s are characterized using MOKE , SQUID and anomalous Hall resistance at different \ncryostat temperature ranging from 10K to 350 K. The current -switching experiments are \nperformed in the spin–orbit torque geometry where the current pulses are injected in plane and the \nmagnetization reversal is detected by measuring the Hall resistance. The full reversal magnetization has \nbeen observed in all samples . Some experimental results could only be explained by the strong sample \nheating effect during the current pulse s injection . We have found that, for a given composition x \nand switching polarity , the devices always reach the same temperature Tswitch (x) before \nswitching independently of the cryostat temperature. Tswitch seems to scale with the Curie \ntemperature of the CoxTb1-x ferrimagnetic alloys. This explain s the evolution of the critical \ncurrent (and critical current density) as a function of the alloy concentration. Future application \ncould take advantages of this heatin g effect which allows reducing the in -plane external field. \nUnexpected double magnetization switching has been observed when the heat generated by \nthe current allow s cross es the compen sation temperature . \n \nI. Introduction \nSpin Orbit Torque switching with perpendicularly magnetized material in Hall bar based devices \noffers a simple and powerful geometry to probe current induced magnetization reversal and \nhad opened a new way to manipulating magnetization at the nanoscale . The underlying physics \nis quite rich and complex including origin of spin-orbit torque (SOT), interfacial effects and \nthermal contributions. Magnetization switching by SOT was first observed in heavy \nmetal/ferromagnetic, HM/FM, ultrathin films [1–3]. The torque is mainly related to the spin Hall \neffect (SHE) [4-10], where the charge current flowing in the heavy metal is converted into a 2 \n vertical spin current due to the large spin -orbit coupling. This spin current is then transferr ed to \nthe FM magnetization, which leads to a torque, namely the spin orbit torque. It has been \ndemonstrated for instance using Pt [4–9], Ta [9–12], W [13–16] as HM and FM layers with \nperpendicular magnetization like C oFeB [7,9–12,14,15] , Co [4–6], CoFeAl [16] or (Co/Ni) [8] \nmultilayers. Interface effects can play a key role in the SOT i n particular, interfacial spin memory \nloss [17] and spin transparency [18] which affects the transmitted spin . Furthermore, \nadditional charge to spin current conversion can also occur due to Edelstein effect [19] in \nRashba [20] and topological insulator interfaces [19]. There is some attempts to unify a \nmodel [21–23] including the aforementioned effects. It might be also an interfacial DMI \n(Dzyaloshinskii -Moriya interaction ) which favors formation of chiral Neel domain wall [24]. It \nwas shown in FM/ HM systems that the reversal of the magnetization occurs first by a magnetic \ndomain nucleation followed by a domain wall prop agation thanks to SHE and iDMI [8,25 –28]. \nThe thermal contribution [6,29] is usually neglected in those experiments. \nFor possible application s the critical switching current need s to be reduced while maintaining a \nsufficient thermal stability. In the literature the critical current density to reverse the \nmagnetization , Jcc, is typically of the order of ~1010 to 1012 A/m2 depending on the applied \ncurrent pulse duration and on the in -plane external magnetic field [4,5,8,30] . Jcc is proportional \nto the m agnetization times the thickness of the FM layer ( Jcc Mt F). Recently , transition metal -\nrare earth TM -RE ferrimagnetic materials start ed to attract large attention for spin -orbitronics \napplications [31–36]. In these ferrimagnetic alloys the net magnetization is given by the sum of \nthe magnetization of the two magnetic sub-lattices (r are earth and transition metal) which are \nantiferromagnetically coupled . The most advantage of ferrimagnetic materials is that its net \nmagnetization M can be tuned by changing its composition or temperature [37]. As a result , a \nmagnetic compensation point with zero magnetization can occur for a certain alloy \nconcentration, xMcomp , or temperature, TMcomp , where the magnetization of both sub -lattices \ncompensates . Moreover TM -RE thin f ilms are characterized by a large bulk magnetic anisotropy \nperpendicular to the film plane (PMA) which make easier to integrate TM -RE with different NM \nmaterials while keeping large thermal stability [38]. Furthermore, the control of magnetization \nswitching using ultrafast femtosecond laser pulse has been demonstrated recently for various \nTM-RE materials [39,40] . Those features are encouraging to combine the control of \nmagnetization by both optical and electrical means. Concerning the Spin Orbit Torque (SOT) \nswitching, reports on experiments with TM-RE alloys claim that the spin -orbit torque efficiency \nreaches a maximum at the magnetic compensation point [31–36], however the critical current is \nnot minimum at this point [33,35,36] . In this study we address the SOT-switching experimen ts \non well characterized //W/Co xTb1-x/Al systems for various concentrations . We demonstrate that \nthermal effects are keys to explain the current induced magnetization reversal in this system . \nWhen the current is injected in the bilayer the Joule heating leads to a large increase of the \nsample temperature . Using systematic SOT measurement s at different temperatures and alloy 3 \n compositions, we establish that for each concentration x the current induced magnetization \nswitching occurs for a unique sample tempe rature Tswitch (x). Tswitch scales with the Curie \ntemperature ( TC) of the alloy. Those new f indings open new rooms to explore combination of \nSOT and thermal contribution towards reducing critical current density to reverse M and \nconsequently low power consu mption applications . In the specific case where Tswitch is close to \nTMcomp an unexpected “double switching ” is observed. \n \nII. Basic characterization \nTo study SOT magnetization switching in RE -TM alloys, a model system composed of CoxTb1-X \nferrimagnetic alloy s deposited on a tungsten heavy metal with high charge to spin conversion \nefficiency [13] was considered . The samples were grown by dc magnetron sputtering on \nthermally oxide Si substrates (Si -SiO 2). The full stacks of the samples are Si -SiO 2//W(3 \nnm)/Co xTb1-x(3.5 nm)/Al(3 nm) with 0.71 x 0.86 . The 3 nm thick Al (naturally oxidized and \npassivated after the deposition ) is used to cap the ferrimagnetic layer. The W and CoTb lay ers \nhave amorphous structure. As described in the introduction , ferrimagnetic alloys like CoTb can \nshow a compensation point at which the Co and Tb moments cancel each other, resulting in \nzero net magnetization . When the net magnetization of the alloy is parallel (resp. antiparallel ) to \nthe magnetization of the Terbium sub -lattice the alloy will be call Terbium rich (resp. Cobalt \nrich) . The samples were characterized by a SQUID -VSM magnetometer and Magneto -optic ally \nKerr effect (MOKE) at room temperature . The SQUID measurements obtained at room \ntemperature are presented in Fig 1a . Magnetization compensation is observed for a \nconcentration x Mcomp =0.77 where the coercivity Hc diverge s and the net saturation \nmagneti zation M s tends to zero . This value is close to the one report ed for bulk and thicker \nCoxTb1-x films [37,41] at room temperature . Additionally to the divergence of Hc, MOKE \nmeasurements show that the Kerr angle rotation changes its sign between Co-rich and Tb-rich \nsamples , which can be explained by the fact that Kerr rotation is mainly sensitive to the Cobalt \nsub-lattice (see for instance fig . S1 in supplementary material [42]). Both SQUID and MOKE \nresults clearly show that all CoTb films studied have a strong out of plane magnetic anisotropy. \nTo study Spin Orbit Torque switching, the stacks were patterned by standard UV lithography \ninto micro -sized Hall crosses with a channel of 2, 4, 10 and 20 m. The results shown ar e \nobtained for a width of 20 m unless other wise specified. Ti(5)/Au(100) ohmic contacts were \ndefined by evaporation deposition and lift-off method on top of W layers . By measuring the \nanomalous Hall resistance RAHE of the hall crosses while sweeping the external perpendicular \nmagnetic field Hz at different temperatures , we could determin e the magnetic compensa tion \ntemperature of the samples . Fig. 1b shows the temperature dependence of Hc for 78% of Co. \nThe coercive field Hc diverge s around 280 K which determines TMcomp for this composition. 4 \n Moreover, we can observe in the insets that the RAHE(Hz) cycle is reversed for Tb -rich (T<280 K) \nand Co -rich( T>280 K) phases , namely change of field switching polarity (Field -SP). The latter is \ndue to the fact the Anomalous hall resistance is sensitive to the cobalt sub -lattice . Van der Pauw \nresistivity measurements leads to a resistivity of W in Si-SiO 2//W(3 nm)MgO(3 nm) of W = 162 \n.cm. Then we could deduce the Co 0.72Tb0.28 resistivity CoTb = 200 .cm which decrease s to \n135 .cm when the Cobalt concentration reaches Co0.86Tb0.14 in accord with previous \nresults [43]. Despite this trend, the a mplitude of RAHE(Hz), RAHE, increases as a function of the \nCo concentration verifying that RAHE is mainly sensitive to the Cobalt sub -lattice (see also Fig \nS2 [42]). \n \nIII. Thermal ly assisted and spin -orbit torque switching \nFig. 2a shows a scheme of a Hall bar along with the convention s used for current injection , \nvoltage probe and directions axes . Typical RAHE(Hz) cycles obtained at room temperature with a \nlow in -plane d c current of 400 A (charge current density of about 2.4 109A/m2 flowing in each \nlayer) for a Tb-rich ( resp. Co-rich) sample is shown in Fig. 2b ( resp. Fig. 2e) . As expected , a \nchange of Field-SP is observed since the alloy net magnetization is parallel to the magnetization \nof the Cobalt sub -lattice in one case and antiparallel in the other . For the same samples the \ncurrent -induced switching cycles are shown in Fig. 2c -d (Tb -rich) and 2f -g (Co -rich) with a n in-\nplane bias field of Hx=100 mT (Fig. 2c and 2f) and -100 mT (Fig. 2d and 2g). The current injection \nwas performed with pulse duration of 100 s using a K6221 source coupled to a K2182 Keithley \nnanovoltmeter. The Hall voltage is measured during the pulse. We have observed the current \ninduced magnetization switching in all the samples for 0.72 x 0.86. The Hall resistance \namplitudes are the same for the current -switching and the field -switching cycles indicating that \nthe reversal of magnetization is fully achieved in both cases . The da ta of the full series are \nshown in Fig. S3 [42]. Sharp current switching are observed and the critical current reduce s \nwhen Hx increases following similar trends that for ferromagnetic materials [30] as shown in \nFig. S4 . Remarkably, we observe a full magnetization reversal even for a n in-plane field Hx as low \nas 2 mT. The role of the in -plane field can be understood as the field to balance the iDMI to \npropagate domain walls which have in -plane magnetization after nucleation of magnetic \ndomains or the field to break the symmetry and to allow for a deterministic switching [8,30] . If \nthe SOT depends on the Co moment , the SOT acts as an effective field HSHE m [24,44] \nwhere m is the magnetic moment and the spin polarization of the spin current Js injected from \nthe W layer into the CoTb layer. is along the y direction in our measurement geometry (it \nchanges between +y and –y when the direction of the injected current is inverted). m changes \nits sign upon the change of the in -plane field direction . Then the sign of the Hall cycle vs current, \nRAHE(i), is reversed when Hx is reversed as observed in Fig 2 c-f. Additionally in ferrimagnetic 5 \n alloys, the effective field HSHE can be reversed if a Co -rich s ample is replaced by a Tb-rich one (a \nschematic is shown in Fig. 2h for Hx>0 and i>0). The identical effect will be observed if the same \nsample is kept and the magnetic compensation temperature is crossed. The fact that the \nsample s which have been identified as Co-rich and Tb -rich at room temperature are showing \nthe same current -switching polarity (Fig 2c and 2f) can only be understood if the so call Tb -rich \nsample has cross ed compensation to become Co -Rich. This compensation crossing is due to the \nJoule heating effect. This assumpt ion was tested by measuring RAHE(Hz) cycles for different \napplied current pulses on the “Tb-rich” sample shown in Figure 3. We observe that for applied \ncurrent s i< 19.5 mA the sign of the cycle demonstrate a “ Tb-rich” nature, however for current \ni>19.5 mA the RAHE(Hz) cycles are reversed and demonstrate a “Co-rich” nature . This is clear \nexperimental evidence that for current close to 19.5 mA the device reach es the sample \ncompensation temperature (TMcomp ~320 K). This demonstrates that the sample is strongly \nheated during the current -switching experiments. The temperature can be determined by the \nresistance value as explained in the next section. In Figure 3 the corresponding temperature is \nshown using color code. We can determine that a temperature of 460 K is reached for 24mA \nwhich is the critical current to switch M (Fig. 2c) . Th is current -switching of 24 mA is then \nobtained for a temperature above TMcomp which explain why the sign of the RAHE(i) cycle is the \none expected for a Co -rich sample. We have performed RAHE(Hz) cycles with intensity current \npulse as high as 34 mA (~525 K) where we can observe that the device shows a ferromagnetic \nhysteresis loop and remain perpendicular ly magneti zed. TC is then higher than 525 K . \n \nIV. Characteristic temperatures of switching \nSince we have addressed the reason of the observed switching polarity several questions are \narising : i) how much are the devices heated when the switching occurs ? ii) Does the \ntemperature at which the switching occurs changes with the initial temperature (temperature \nat which the experiment is carried out, T cryostat ) ? iii) How d oes the switching current and \nswitching temperature depend on composition? And iv ) What is the physical meaning of this \nswitching temperature: Angular compensation temperature T Acomp ? In order to address all those \nquestions we have performed a series of temperature dependence experiments for various \nsamples. \nFig. 4 a shows the RAHE(ipulse) cycles for Hx>0 at different cryostat temperature for W/Co0.73Tb0.27 \n(Tb-rich at room temperature). We have observed the Down -Up current -switching polarity at \n300 K . For 150 K Tcryostat 250 K a double current switching loop is observed. This type of \ndouble current switching can be explain ed when the switching temperat ure is close to TMcomp \nand its origin will be discussed later in the paper (Fig. 4a show s only the case of 150 K for \nclarity ). For 10 K T 100 K we obser ve only Down -Up current -switching polarity (we didn’t 6 \n increase too much the pulse current to avoid burn ing the device). One can calibrated the real \nsample temperature at different pulse -current performing the following protocol: i) measuring \nthe resistance of the current channel Rchannel (ipulse) as function of pulse current intensity as \nshown in Fig. 4b for different cryostat temperatures, and ii) measuring the temperature \ndependence of the current channel Rchannel (T) as shown in F ig 4c (for which we use a very low dc \nbias current of only 400 A). Interesting ly, we observe that for Co -rich current -switching polarity \n(Down -Up) the device reaches the same resistance (1.373 k ) and consequently the same \nswitching temperature Tswitch = 435 K 25 K for this W/Co 0.73Tb0.27. We note that the resistance \ndecreases when T increases which is a feature and confirmation of amorphous materials [45]. \nWe have performed the same protocol for various compositions and different devices. An \nexam ple for Co -rich sample at room temperature is shown in Fig. 4 d-f (W/Co 0.79Tb0.21) where \nwe also observe that the critical current heat s up the device to the same channel resistance (Fig. \n4e), so the same T switch (~485 K for this Co xTb1-x sample) irrespective of the initial temperature . \nMoreover , on this particular sample TMcomp is about 500 K and we observe no change of C urrent -\nSP even for T as low as 10 K which is well below its TMcomp . \nAdditionally to the characteristic Tswitch we have just disc ussed , one can also investigate the \ntemperature dependence of the critical current as shown in Fig . 5a for a Co0.78Tb0.22 sample ( Co-\nrich at room temperature ). The extrapolation of the linear dependence to zero current is \ndefined as T*. In Fig. 5b is found out that T*~470 K for W/ Co 0.78Tb0.22. \nFig. 6a show s RAHE(ipulse) for Co 0.72Tb0.28 performed for a cryostat temperature of 150 K to \nhighl ight the observation of the two switching. At lower current (37.5 mA) , i.e. lower Joule \nheating effect, we observed Up-Down current -switching polarity while the second one which \noccurs at higher current (43 mA) is Down -Up. This can be understood considering that to \nachieve the first switching the device reaches a temperature below its TMcomp so the sample is \nstill Tb -rich and the Up -Down current -switching polarity observed is as expected (i.e Fig 2h) . If \nwe continue increas ing the intensity of the applied in -plane pulse current we overcome TMcomp \nand the n the perpendicular component of effective torque field HSHE now changes its sign as \ndiscussed previous ly. Consequently, the second observed switching agree well for Co -rich phase \n(Down -Up). In Fig. 6b is shown the temperature dependence of both switching current s. As \ndiscussed, the first (second) switching agrees with a Tb -rich (Co -rich) current -switching polarity \nand happens for TTMcomp ). The linear extrapolation of both switching currents \nroughly tends to 350 K (Tb-rich switching ) and T*~455 K (Co-rich switching ). The value of T* \nseems to be slightly higher than Tswitch (~435 K 20 K as determined in Fig 4c). \n \n 7 \n V. T-x switching phase diagram and conclusions \nIn figure 7a the different characteristic temperature s of our W/CoTb systems can be plotted in \nthe (T,xCo) phase diagram. The determine d TMcomp decreases linearly with the Co concentration \nas reported for bulk CoTb and thick CoGd films (300 nm) [41,46] . However Tswitch and T* \nincrease l inearly with the Co-concentration and scale with the Curie temperature Tc thus \ndepend ing on composition , and independent of initial temperature. It is remarkable that the \nTswitch and T* are nearly the same, indicating that, to achieve the switching, one has to reach a \nspecific temperature. The three first questions in section IV are answered . Now let’s discuss the \nphysical meaning of these switching temperatures. It is clear that for Co -Current SP the \ntemperature of switching is above TMcomp and below Tc. Fig. 7a also shows Tc in bulk CoTb after \nHans et al. [41]. The angular composition temperature , TAcomp , scales with TMco mp. Indeed, \ntypically TAcomp ~(TMcomp +30 K) for Co0.775Gd 0.225 thick films (300 nm) [46]. This is explained by \nthe relationship between the angular moment L, magnetic moment and the gyromagnetic ratio \n or Landé g -factor ( LTM, RE =MTM, RE /TM, RE and =gB/hb). Therefore in CoTb it is expected that the \ntrends of TAcomp and TMcomp are similar and they decreas e with increasing Co concentration . \nHowever Tswitch increases with Co concentration which indicates that Tswitch is not scaling with \nTAcomp or TMcomp. For sake of comparison we plotted the switching -current for experiment \nperformed at room temperature together with the temperature increase T = TswitchTcryostat , \ndue to Joule heating effect, Fig. 7b . We can observe that the temperature is increased between \n100 K and 300 K. This variation will increase when we reduce Tcryostat . Considering that resistivity \nof both, CoTb and W layers, change similarly with temperature and using the resistivity \nmeasured at room temperature we can estimate the critical current density JCC flowing on each \nlayer as displayed in Fig. 7c. We observe that Jcc on W is reduce d by a factor of ~2 wh ile varying \nthe composition of CoTb . We observe the minimum of Jcc at the lower Co concentration \nmeasured. This c an be explained by the fact that TC and Tswitch decrease with de creas ing Co-\nconcentration but a relation ship between Tswitch and TC will need additional studies . \nConclusions \nIn conclusion w e have fully characterized the current -induced switching experiment in a series \nof //W(3nm)/Co xTb1-x(3.5nm)/AlO x(3nm) samples. In addition to the SOT effect we demonstrate \na strong thermal contribut ion to achieve the magnetization reversal. For the Co -rich current -\nswitching polarity the device needs to reach the same temperature Tswitch to achieve the \nswitching. This Tswitch increase s with Co -concentration which then scale with Curie temperature \nTc. It is then unlikely that Tswitch corresponds to angular momentum compensation temperature \n(which scales with TMcomp decreasing with Co -concentration) . Those results highl ight the \nimportance of considering thermal contributions in SOT switching experiment s and the fact that \nthe spin Hall angle determination might be overestimated when thermal contribution is 8 \n neglected . The use of resistive W layer i ncrease the heating of the device, reducing strongly the \nexternal in -plane needed to assist the SOT . Those results are important for the full \nunderstanding of current -induced magne tization switching and may lead the way to new \ntechnological applications taking advantages of the rather strong heating \nAcknowledgements \nThis work was supported partly by the french PIA project “Lorraine Université d’Excellence”, \nreference ANR -15-IDEX -04-LUE. by the ANR -NSF Project, ANR -13-IS04 -0008 - 01, “COMAG” by \nthe ANR -Labcom Project LSTNM, Experiments were performed using equipment fro m the \nTUBE —Daum funded by FEDER (EU), ANR, the Region Lorraine and Grand Nancy. We thank J. \nSampaio and S. Petit-Watelot for fruitful discussions. \n \n \nReferences \n[1] J. E. Hirsch, “Spin Hall Effect” Phys. Rev. Lett. 83, 1834 (1999). \n[2] A. Hoffmann, “Spin Hall Effects in Metals” IEEE Trans. Magn. 49, 5172 (2013). \n[3] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, “Spin Hall effects” Rev. \nMod. Phys. 87, 1213 (2015). \n[4] I. M. Miron, K. Garello, G. Gaudin, P. -J. Zermatten, M. V Costache, S. Auffret, S. Bandiera, B. \nRodmacq, A. Schuhl, and P. Gambardella, “Perpendicular switching of a single ferromagnetic layer \ninduced by in -plane current injection.” Nature 476, 189 (2011). \n[5] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and R. a. Buhrman, “Current -induced switching of \nperpendicularly magnetized magnetic layers using spin torque from the spin hall effect” Phys. \nRev. Lett. 109, 1 (2012). \n[6] C. Bi, L. Huang, S. Long, Q. Liu, Z. Yao, L. Li, Z. Huo, L. Pan, and M. Liu, “Thermally assisted \nmagnetic switching of a single perpendicularly magnetized layer induced by an in -plane current” \nAppl. Phys. Lett. 105, 1 (2014). \n[7] X. Qiu, K. Narayanapillai, Y. Wu, P. Deorani, D. -H. Yang, W. Noh, J. P ark, K. -J. Lee, H. Lee, and H. \nYang, “Spin –orbit -torque engineering via oxygen manipulation” Nat. Nanotechnol. 10, 333 (2015). \n[8] J. C. Rojas -Sánchez, P. Laczkowski, J. Sampaio, S. Collin, K. Bouzehouane, N. Reyren, H. Jaffrès, A. \nMougin, and J. M. George , “Perpendicular magnetization reversal in Pt/[Co/Ni]3/Al multilayers via \nthe spin Hall effect of Pt” Appl. Phys. Lett. 108, 82406 (2016). \n[9] D. Wu, G. Yu, Q. Shao, X. Li, H. Wu, K. L. Wong, Z. Zhang, X. Han, P. Khalili Amiri, and K. L. Wang, \n“In-plane cu rrent -driven spin -orbit torque switching in perpendicularly magnetized films with \nenhanced thermal tolerance” Appl. Phys. Lett. 108, 212406 (2016). 9 \n [10] L. Liu, C. -F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, “Spin -Torque Switching with the \nGiant Spin Hall Effect of Tantalum” Science (80 -. ). 336, 555 (2012). \n[11] C. Zhang, S. Fukami, H. Sato, F. Matsukura, and H. Ohno, “Spin -orbit torque induced \nmagnetization switching in nano -scale Ta / CoFeB / MgO” Applied 107, 12401 (2015). \n[12] Y. M. Hung, L. Rehm, G. Wolf, and A. D. Kent, “Quasistatic and Pulsed Current -Induced Switching \nwith Spin -Orbit Torques in Ultrathin Films with Perpendicular Magnetic Anisotropy” IEEE Magn. \nLett. 6, (2015). \n[13] C.-F. Pai, L. Liu, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, “Spin transfer torque devices \nutilizing the giant spin Hall effect of tungsten” Appl. Phys. Lett. 101, 122404 (2012). \n[14] Q. Hao and G. Xiao, “Giant Spin Hall Effect and Switching Induced by Spin -Transfer Torque in a \nW/CoFeB/MgO structure with perpendicular magnetic anisotropy” Phys. Rev. Appl. 3, 34009 \n(2015). \n[15] S. Cho, S. -H. C. Baek, K. -D. Lee, Y. Jo, and B. -G. Park, “Large spin Hall magnetoresistance and its \ncorrelation to the spin -orbit torque in W/CoFeB/MgO structures.” Sci. Rep. 5, 14668 (2015). \n[16] M. S. Gabor, T. Petrisor, R. B. Mos, A. Mesaros, M. Nasui, M. Belmeguenai, F. Zighem, and C. \nTiusan, “Spin –orbit torques and magnetization switching in W/Co 2 FeAl/MgO structures” J. Phys. \nD. Appl. Phys. 49, 365003 (2016). \n[17] J.-C. Rojas -Sánchez, N. Reyren, P. Laczkowski, W. Savero, J. -P. Attané, C. Deranlot, M. Jamet, J. -M. \nGeorge, L. Vila, and H. Jaffrès, “Spin Pumping and Inverse Spin Hall Effect in Platinum: The \nEssential Role of Spin -Memory Loss at Metallic Interfaces” Phys. Rev. Lett. 112, 106602 (2014). \n[18] W. Zhang, W. Han, X. Jiang, S. -H. Yang, and S. S. P. Parkin, “Role of transparency of platinum –\nferromagnet interfaces in determining the intrinsic magnitude of the spin Hall effect” Nat. Phys. \n11, 496 (2015). \n[19] J. C. Rojas -Sánchez, S. Oyarzún, Y. Fu, A. Marty, C. Vergnaud, S. Gambarelli, L. Vila, M. Jamet, Y. \nOhtsubo, A. Taleb -Ibrahimi, P. Le Fèvre, F. Bertran, N. Reyren, J. M. George, and A. Fert, “Spin to \nCharge Conversion at Room Temperature by Spin Pumping in to a New Type of Topological \nInsulator: α -Sn Films” Phys. Rev. Lett. 116, 96602 (2016). \n[20] J. C. Rojas Sánchez, L. Vila, G. Desfonds, S. Gambarelli, J. P. Attané, J. M. De Teresa, C. Magén, and \nA. Fert, “Spin -to-charge conversion using Rashba coupling a t the interface between non -magnetic \nmaterials” Nat. Commun. 4, 2944 (2013). \n[21] P. M. Haney, H. -W. Lee, K. -J. Lee, A. Manchon, and M. D. Stiles, “Current -induced torques and \ninterfacial spin -orbit coupling” Phys. Rev. B 88, 214417 (2013). \n[22] V. P. Amin and M. D. Stiles, “Spin transport at interfaces with spin -orbit coupling: \nPhenomenology” Phys. Rev. B 94, 104420 (2016). \n[23] V. P. Amin and M. D. Stiles, “Spin transport at interfaces with spin -orbit coupling: Formalism” \nPhys. Rev. B 94, 104419 (2016). \n[24] A. Thiaville, S. Rohart, É. Jué, V. Cros, and A. Fert, “Dynamics of Dzyaloshinskii domain walls in \nultrathin magnetic films” EPL (Europhysics Lett. 100, 57002 (2012). 10 \n [25] O. J. Lee, L. Q. Liu, C. F. Pai, Y. Li, H. W. Tseng, P. G. Gowtham, J. P. Park, D. C. Ralph, and R. a. \nBuhrman, “Central role of domain wall depinning for perpendicular magnetization switching \ndriven by spin torque from the spin Hall effect” Phys. Rev. B 89, 24418 (2014). \n[26] C. F. Pai, M. Mann, A. J. Tan, and G. S. D. Beac h, “Determination of spin torque efficiencies in \nheterostructures with perpendicular magnetic anisotropy” Phys. Rev. B 93, 144409 (2016). \n[27] N. Mikuszeit, O. Boulle, I. M. Miron, K. Garello, P. Gambardella, G. Gaudin, and L. D. Buda -\nPrejbeanu, “Spin -orbit torque driven chiral magnetization reversal in ultrathin nanostructures” \nPhys. Rev. B 92, 144424 (2015). \n[28] M. Baumgartner, K. Garello, J. Mendil, C. O. Avci, E. Grimaldi, C. Murer, J. Feng, M. Gabureac, C. \nStamm, Y. Acremann, S. Finizio, S. Wintz, J. Raabe, and P. Gambardella, “dynamics driven by spin \n– orbit torques” Nat. Nanotechnol. 12, 980 (2017). \n[29] K.-S. Lee, S. -W. Lee, B. -C. Min, and K. -J. Lee, “Thermally activated switching of perpendicular \nmagnet by spin -orbit spin torque” Appl. Phys. Lett. 104, 72413 (2014). \n[30] K.-S. K. -J. Lee, S. -W. Lee, B. -C. Min, and K. -S. K. -J. Lee, “Threshold current for switching of a \nperpendicular magnetic layer induced by spin Hall effect” Appl. Phys. Lett. 102, 112410 (2013). \n[31] N. Roschewsky, T. Matsumura, S. C heema, F. Hellman, T. Kato, S. Iwata, and S. Salahuddin, “Spin -\nOrbit Torques in ferrimagnetic GdFeCo Alloys” Appl. Phys. Lett. 109, 112403 (2016). \n[32] K. Ueda, M. Mann, C. Pai, A. Tan, G. S. D. Beach, K. Ueda, M. Mann, C. Pai, A. Tan, and G. S. D. \nBeach, “Spin -orbit torques in Ta / TbxCo100 -x ferrimagnetic alloy films with bulk perpendicular \nmagnetic anisotropy Spin -orbit torques in Ta / Tb x Co 100 -x ferrimagnetic alloy films with bulk \nperpendicular magnetic anisotropy” Appl. Phys. Lett. 109, 232403 (2016 ). \n[33] J. Finley and L. Liu, “Spin -Orbit Torque Efficiency in Compensated Ferrimagnetic Cobalt -Terbium \nAlloys” Phys. Rev. Appl. 54001 , 1 (2016). \n[34] N. Roschewsky, C. Lambert, and S. Salahuddin, “Spin -orbit torque switching of ultralarge -\nthickness ferrimagnetic GdFeCo” Phys. Rev. B 96, 64406 (2017). \n[35] W. S. Ham, S. Kim, D. Kim, K. Kim, T. Okuno, H. Yoshikawa, A. Tsukamoto, T. Moriyama, and T. \nOno, “ Temperature dependence of spin -orbit effective fields in Pt / GdFeCo bilayers Temperature \ndependence of spin -orbit effective fields in Pt / GdFeCo bilayers” Appl. Phys. Lett. 110, 242405 \n(2017). \n[36] R. Mishra, J. Yu, X. Qiu, M. Motapothula, T. Venkatesan, and H. Yang, “Anomalous Current -\nInduced Spin Torques in Ferrimagnets near Compensation” Phys. Rev. Lett. 118, 167201 (2017). \n[37] M. Gottwald, M. Hehn, F. Montaigne, D. Lacour, G. Lengaigne, S. Suire, and S. Mangin, \n“Magnetoresistive effects in perpendicu larly magnetized Tb -Co alloy based thin films and spin \nvalves” J. Appl. Phys. 111, (2012). \n[38] S. Mangin, T. Hauet, P. Fischer, D. H. Kim, J. B. Kortright, K. Chesnel, E. Arenholz, and E. E. \nFullerton, “Influence of interface exchange coupling in perpendi cular anisotropy [Pt/Co]50/TbFe \nbilayers” Phys. Rev. B 78, 24424 (2008). \n[39] C.-H. Lambert, S. Mangin, B. S. D. C. S. Varaprasad, Y. K. Takahashi, M. Hehn, M. Cinchetti, G. 11 \n Malinowski, K. Hono, Y. Fainman, M. Aeschlimann, and E. E. Fullerton, “All -optical control of \nferromagnetic thin films and nanostructures.” Science 345, 1337 (2014). \n[40] S. Mangin, M. Gottwald, C. -H. Lambert, D. Steil, V. Uhlíř, L. Pang, M. Hehn, S. Alebrand, M. \nCinchetti, G. Malinowski, Y. Fainman, M. Aeschlimann, and E. E. Fullerton, “Engineered materials \nfor all -optical helicity -dependent magnetic switching.” Nat. Mater. 13, 286 (2014). \n[41] P. Hansen, C. Clausen, G. Much, M. Rosenkranz, and K. Witter, “Magnetic and magneto -optical \nproperties of rare -earth transition -metal alloys con taining Gd, Tb, Fe, Co” 756, (1989). \n[42] T. H. Pham, S. G. Je, P. Vallobra, F. Thibaud, D. Lacour, G. Malinowski, M. C. Cyrille, G. Gaudin, O. \nBoulle, M. Hehn, J. C. Rojas -Sanchez, and S. Mangin, “Thermal contribution to the spin -orbit \ntorque in metallic/ ferrimagnetic systems” Suppl. Inf. (2017). \n[43] T. W. Kim and R. J. Gambino, “Composition dependence of the Hall effect in amorphous \nComposition dependence of the Hall effect in amorphous Tb x Co 1 À x thin films” J. Appl. Phys. \n87, 1869 (2000). \n[44] a. V. Khvalkovskiy, V. Cros, D. Apalkov, V. Nikitin, M. Krounbi, K. a. Zvezdin, A. Anane, J. Grollier, \nand A. Fert, “Matching domain -wall configuration and spin -orbit torques for efficient domain -wall \nmotion” Phys. Rev. B 87, 20402 (2013). \n[45] A. Fert, R. Aso moza, A. Fert, R. Asomoza, S. Universite, and P. O. France, “Transport properties of \nmagnetic amorphous alloys Transport properties of magnetic amorphous alloys” 1886 , (1979). \n[46] M. Binder, a. Weber, O. Mosendz, G. Woltersdorf, M. Izquierdo, I. Neudecker, J. Dahn, T. \nHatchard, J. -U. Thiele, C. Back, and M. Scheinfein, “Magnetization dynamics of the ferrimagnet \nCoGd near the compensation of magnetization and angular momentum” Phys. Rev. B 74, 134404 \n(2006). \n \n \n 12 \n \n \nFigure 1. //W(3)/Co xTb1-x (3.5) /Al(3) : a) Coercive field Hc and saturation magnetization Ms obtained by \nSQUID measurements vs. cobalt concentration at room temperature. Magnetic compensation is \nobserved for xMcomp ~0.77. b) Temperature dependence of coercitivity on Hall bar for x=0.78 showing that \nTMcomp ~280 K for x=0.78. Insets show Hall resistance cycle for a temperature below (left) and above \n(right) TMcomp . The change of field - switching polarity (Field -SP) evi dences that RAHE is mainly sensitive to \nthe magnetization of the Cobalt sub -lattice. \n \n13 \n \n \nFigure 2. Anomalous Hall effect and Current -induced magnetization reversal at room temperature on Si -\nSiO 2//W(3)/Co xTb1-x (3.5) /Al(3). a) Scheme of the Hall bar along with geometry used. The spin \npolarization is along y axes. (b and g ) Sweeping perpendicular field (H ||z) with a low dc bias of 400 A. \n(c,d,e and f ) Sweeping in -plane current (i pulse ||x) with an in -plane field Hx=±100 mT .The width of the \nchannel current is 20 m. The current switching polarity (current -SP) for Hx>0 is Down -Up for Co -rich and \nTb-rich samples (c,f). The Current -SP in (c) is opposite than predicted for Tb -rich samples as shown in the \nschematic in (h). The p erpendicular effective torque field HSHE is proportional to m and should lead to \na change of the sign in the RAHE(i) cycle when changing from Co -rich to Tb -rich phase. \n \n14 \n \n \nFigure 3. R AHE(Hz, ipulse) cycles on Si -SiO 2//W(3)/Co 0.76Tb0.24(3.5)/Al(3) Hall bar measured at room \ntemperature. The cycles are vertically offset for clarity. It is observed that for i < 19.5 mA the cycle has a \nsignature corresponding to Tb -rich phase according to our convention. However for i> 19.5 mA the cycle \nchanges their sign and now corresponds to Co -rich phase. It is an evidence of the Joule heating effect \nwhen high pulse current is applied. 19.5 mA roughly corresponds to TMcomp . The critical current for this \ndevice is about 24 mA. Thus during the electrical switching T device>TMcomp >300 K for this Co 0.76 system. \n \n15 \n \n \n \nFig4. a) RAHE(ipulse) at different cryostat temperatures. Cycles for x=0.73 (Tb -rich at room temperature ). b) \nRChannel (ipulse) and c) RChannel (T) at different cryostat temperatures. The vertical dashed line points out the \ncritical current to reverse M. It is observed that independently of the initial temperature, the device \nalways reaches the same value of longitudinal resistance (1373 ) which means it reaches the same \ntemperature . The linear extrapolation of RChannel (T) allows us to know the temperature corresponding to \nthe current -induced magnetization reversal . Such a temperature is defined as Tswitch. RChannel (T) is \nperformed with a low bias current of 400 A. For Co 0.73 we found that Tswitch= 435 K 25 K. d) RAHE(ipulse) \ncycles for x=0.79 (Co -rich at room temperature) at different cryostat temperatures. e) RChannel (ipulse) and f) \nRChannel (T) at different cryostat temperatures. It is also observed that independently of the initial \ntemperature, the device always reaches the same resistance , thus is the same temperature . In this case \nit corresponds to 1291 and Tswitch ~ 485 K. \n \n16 \n \n \nFigure 5 . a) RAHE(ipulse) at different cryostat temperatures. Cycles for x=0.78 (Co -rich at r oom \ntemperature ). b) The critical current to reverse M increase s linearly when T decrease and saturate for T< \n50 K. The extrapolation of th e linear behavior at higher temperature for zero current is defined as T*. \n \n17 \n \n \nFig6. a) RAHE(ipulse) at 150K for x=0.73 (Tb -rich at room temperature ). There are two switching s: i) at lower \ncurrent it agrees with a Tb-rich switching polarity (Up-Down) . ii) The reversal with hi gher current agrees \nwith a Co -rich switching polarity (Down -Up). b) Temperature dependence o f the critical current s for this \ncomposition. TMcomp would be between 350 K and 455 K. \n \n18 \n \n \nFigure 7 . a) Characteristic temperatures as function of Co concentration: T*, Tswitch and TMcomp \nindependently measured in Hall bar patterned devices (lines are guides for the eyes). It is observed that \nTMcomp follows the same behavior reported for bulk Co xTb1-x alloys. T* and Tswitch have the same trend \nthan that of the Curie temperature Tc. The gre en dashed line stands for Tc in bulk CoTb after Hans et \nal. [41]. b) The total critical current injected to reverse M when the experiment is performed at room \ntemperature (= Tcryostat ), and the variation of temperature TswitchTcryostat to reach the switching. c) The \ncritical current density, calculated from b, flowing in W and CoTb layers, respectively. \n \n1 \n Supplementary Material \nThermal contribution to the spin -orbit torque in metallic/ferrimagnetic systems \nThai Ha Pham1, S.-G. Je1,2, P. Vallobra1, T. Fache1, D. Lacour1, G. Malinowski1, M. C. Cyrille3, G. Gaudin2, O. \nBoulle2, M. Hehn1, J.-C. Rojas -Sánchez1* and S. Mangin1 \n1 . Institut Jean Lamour, CNRS UMR 7198, Université de Lorraine, F-54011 Nancy, France \n2 . CNRS, SPINTEC, F -38000 Grenoble \n3. Leti, technology research institute, CEA, F -38000 Grenoble \n*juan -carlos.rojas -sanchez@univ -lorraine.fr \n \nS1- Magneto -optically Kerr effect measurements at room temperature \nFigure S1 show s the coercitivity obtained by MOKE of W/Co xTb1-x thin film as a function of x (the Co-\nconcentration ). The results show that coercive field Hc diverges about Co -concentration of 7 7 % in \nagreement with SQUID results (Fig. 1a). The insets highlight the opposite sign of Kerr angle rotation \nwhen the CoTb alloy change from Tb -rich to Co -rich phase. \n \n \nFigure S1 . //W(3)/Co xTb1-x (3.5) /Al(3) : Coercive field Hc obtained by MOKE measurements at room \ntemperature as a function of the Cobalt concentration . Insets show raw data of MOKE cycles for a Tb -\nrich (left) and Co -rich (right) samples. \n \n \n2 \n S2- Hall resistance amplitude | RAHE|, channel resistance Rchannel , and Co xTb1-x \nresistivity CoxTb1 -x at room temperature \nFigure S2 present the evolution of the Hall resistance amplitude |RAHE|, the channel resistance Rchannel , \nand the CoxTb1-x resistivity CoTb at room temperature as a function of the Co -concentration . Despite the \ndecreas e of the channel resistance with increasing Co -concentration, we observe that | RAHE| increase s \nwith Co-concentration . Thus corroborate that the Hall resistance is mostly sensitive to Co magnetic \nmoment. The W resistivity, W = 162 .cm, is of similar order than CoTb alloys. \n \n \nFigure S2 . Si-SiO 2//W(3)/Co xTb1-x (3.5) /Al(3) : The change of Hall resistance amplitude |RAHE|, channel \nresistance Rchannel , and Co xTb1-x resistivity as a function of the Cobalt concentration at room temperature . \nThe vertical blue dashed line points correspond to the magnetic compensation point at room \ntemperature. The horizontal red dashed line shows the value of W resistivity (ρ W). \n \n \n \n3 \n S3- Current -switching in the full W/CoxTb1 -x series for different in -plane field at \nroom temperature \nFigure S3 (S4) presents the c urrent -induced magnetization reversal performed at room temperature on \nSi-SiO 2//W(3 nm)/Co xTb1-x (3.5 nm) /AlOx(3 nm) with an in -plane field Hx=100 mT (5 mT). \n \n \nFigure S3 . Current -induced magnetizati on reversal at room temperature : Hall resistance as a function of \nthe injected current measured for various ipulse Si-SiO 2//W(3 nm)/Co xTb1-x (3.5 nm) /AlOx(3 nm), with an \nin-plane field of Hx= 100 mT. \n4 \n \nFigure S4 . Current -induced magnetization reversal at room temperature : Hall resistance as a function of \nthe injected current measured for various ipulse Si-SiO 2//W(3 nm)/Co xTb1-x (3.5 nm) /AlOx(3 nm), with an \nin-plane field of Hx= 5 mT. \n \n \n5 \n S4- Power consumption in the W/Co xTb1-x Hall bar with 20 m of width at room \ntemperature \nFigure S 5 shows the electrical power consumption to revers e the magnetization i n Hall bar of length L= \n100 m and width of w = 20 m at room temperature on various Si-SiO 2//W(3 nm)/Co xTb1-x (3.5 nm) \n/AlOx (3nm) with an in -plane field of 100 mT. \n \n \nFigure S5 . Si-SiO 2//W(3)/Co xTb1-x (3.5) /AlOx(3): The electrical power consumption to switch the \nmagnetization at room temperature of a Si-SiO 2//W(3)/Co xTb1-x (3.5) /AlOx(3) hall bar as a function of \nCobalt concentration. The current channe l dimensions are length L= 100 m, and width of w = 20 m. \n \n6 \n S5- Hx-I switching phase diagram in the W/Co xTb1-x \nFigure S6 (S7) presents a 2D plot summarizing the c urrent –switching cycles performed under different \nexternal in -plane field Hx (phase di agram ) at room temperature for a channel width of w= 20 m (10 \nm). Figure S6 are the results obtained for Si-SiO 2//W(3 nm)/Co 0.86Tb14 (3.5 nm) /AlOx(3 nm) and fig. S7 \nfor Si-SiO 2//W(3 nm)/Co 0.84Tb0.16 (3.5 nm) /AlOx(3 nm). \n \n \n \nFigure S6 . 2D-plot of c urrent –switching cycles performed at room temperature on Si -SiO 2//W(3 \nnm)/Co 0.86Tb14 (3.5 nm) /AlOx(3 nm) for a channel width of w = 20 m. The R(ipulse, H x) cycles were carried \nout with different applied in plane field between –3 kG and + 3 kG. The red (blue) color region stand for \nUp (Down) magnetic configuration according the schematic Hall bar shown in Fig. 2a. \n \n7 \n \nFigure S 7. 2D-plot of c urrent –switching cycles performed at room temperature on Si -SiO 2//W(3 \nnm)/Co 0.84Tb16 (3.5 nm) /AlOx(3 nm) for a channel width of w = 10 m. The R(ipulse, H x) cycles were carried \nout with different applied in plane field between –6.5 kG and +6.5 kG. The red (blue) color region stand \nfor Up (Down) magnetic configuration according the schematic Hall bar shown in Fig. 2a. \n \n" }, { "title": "0811.2118v1.Selection_rules_for_Single_Chain_Magnet_behavior_in_non_collinear_Ising_systems.pdf", "content": "arXiv:0811.2118v1 [cond-mat.mtrl-sci] 13 Nov 2008Selection rules for Single-Chain-Magnet behavior in non-c ollinear Ising systems\nAlessandro Vindigni1∗and Maria Gloria Pini2\n1Laboratorium f¨ ur Festk¨ orperphysik, Eidgen¨ ossische Te chnische Hochschule Z¨ urich, CH-8093 Z¨ urich, Switzerlan d\n2Istituto dei Sistemi Complessi, Consiglio Nazionale delle Ricerche,\nVia Madonna del Piano 10, I-50019 Sesto Fiorentino (FI), Ita ly\n(Dated: November 2, 2018)\nThe magnetic behavior of molecular Single-Chain Magnets is investigated in the framework of a\none-dimensional Ising model with single spin-flip Glauber d ynamics. Opportune modifications to\nthe original theory are required in order to account for reci procal non-collinearity of local anisotropy\naxes and the crystallographic (laboratory) frame. The exte nsion of Glauber’s theory to the case of\na collinear Ising ferrimagnetic chain is also discussed. Wi thin this formalism, both the dynamics\nof magnetization reversal in zero field and the response of th e system to a weak magnetic field,\noscillating in time, are studied. Depending on the geometry , selection rules are found for the\noccurrence of slow relaxation of the magnetization at low te mperatures, as well as for resonant\nbehavior of the a.c.susceptibility as a function of temperature at low frequenc ies. The present\ntheoryapplies successfully tosome real systems, namely Mn -, Dy-, andCo-based molecular magnetic\nchains, showing that Single-Chain-Magnet behavior is not o nly a feature of collinear ferro- and\nferrimagnetic, but also of canted antiferromagnetic chain s.\nPACS numbers: 75.10.-b, 75.10.Pq, 75.50.Xx, 75.60.Jk\nI. INTRODUCTION\nSlow dynamics of the magnetization reversal is a cru-\ncial requirement for potential applications of Single-\nChain Magnets (SCM’s)1,2,3, and nanowires in general,\nin magnetic-memory manufacture. For nanowires with\na biaxial anisotropy, provided that their length is much\ngreater than the cross section diameter but smaller than\nexchange length, this phenomenon is governed by ther-\nmal nucleation and propagation of soliton-antisoliton\npairs; the associated characteristic time is expected to\nfollow an Arrhenius law4,5. For genuine one dimensional\n(1D) Ising systems with single spin-flip stochastic dy-\nnamics, a slow relaxation of the magnetization was first\npredicted by Glauber6in 1963. Through Glauber’s ap-\nproach,manyphysicalsystemswereinvestigated,ranging\nfrom dielectrics7,8,9to polymers7,10,11. More fundamen-\ntally, this model has been employed to justify the use\noftheKohlrauch-Williams-Wattsfunction10,12(stretched\nexponential) to fit the relaxation of generalized 1D spin\nsystems. Also the universality issue of the dynamic crit-\nical exponent13,14,15,16,17of the Ising model18, as well as\nstrongly out-of-equilibrium processes (magnetization re-\nversal19, facilitated dynamics19, etc.) have been studied\nmoving from the basic ideas proposed by Glauber.\nIn this paper, single spin-flip Glauber dynamics is\nused to investigate theoretically the slow dynamics of\nthe magnetization reversal in molecular magnetic sys-\ntems. In particular, we extend Glauber’s theory6to\nthe Ising collinear ferrimagnetic chain, as well as to the\ncase of a chain in which reciprocal non-collinearity of lo-\ncal anisotropy axes and the crystallographic(laboratory)\nframe is encountered. Such extensions are motivated by\nthefact that(i) inmolecular-basedrealizationsofSCM’s,\nantiferromagnetic coupling typically has a larger inten-\nsity than the ferromagnetic one; in fact, the overlappingof magnetic orbitals, which implies antiferromagnetic ex-\nchange interaction between neighboring spins, can be\nmore easily obtained than the orthogonality condition,\nleadingtoferromagnetism20,21,22; (ii)non-collinearitybe-\ntweenlocalanisotropyaxesandthecrystallographic(lab-\noratory) frame takes place quite often in molecular spin\nchains. Besides magnetization reversal, the dynamic re-\nsponse of the system to a weak magnetic field, oscillating\nin time at frequency ω, is also studied. Depending on the\nspecific experimental geometry, selection rules are found\nfor the occurrence of resonant behavior of the a.c.sus-\nceptibility as a function of temperature (stochastic reso-\nnance) at low frequencies, as well as for slow relaxation\nof the magnetization in zero field at low temperatures.\nThe paper is organized as follows. In Sect. II we ex-\ntend Glauber’s theory6, originally formulated for a chain\nofidenticalandcollinearspins, tothemoregeneralmodel\nof a chain with non-collinear spins, possibly with Land´ e\nfactors that vary from site to site. In Sect. III we use\ntwo different theoretical methods (the Generating Func-\ntions approach and the Fourier Transform approach) to\ninvestigate the relaxation of the magnetization after re-\nmoval of an external static magnetic field, starting from\ntwo different initial conditions: fully saturated or par-\ntially saturated. In Sect. IV we calculate, in a linear ap-\nproximation, the magnetic response of the system to an\noscillatingmagnetic field. Fora chain of Nspins, the a.c.\nsusceptibility is expressed as the superposition of Ncon-\ntributions, eachcharacterizedby itstime scale; througha\nfew simple examples, we show that, depending on the ge-\nometryof the system ( i.e., the relativeorientationsofthe\nlocal easy anisotropy axes and of the applied field), dif-\nferent time scalescan be selected, possibly giving rise, for\nlow frequencies, to a resonant peak in the temperature-\ndependence of the complex magnetic susceptibility. In\nSect. V we show that the theory provides a satisfactory2\naccount for the SCM behavior experimentally observed\nin some magnetic molecular chain compounds, charac-\nterized by dominant antiferromagnetic exchange interac-\ntionsandnon-collinearitybetween spins. Finally, inSect.\nVI, the conclusions are drawn and possible forthcoming\napplications are also discussed.\nII. THE NON-COLLINEAR ISING-GLAUBER\nMODEL\nIn a celebrated paper6, Glauber introduced, in the\nusual 1D Ising model18, a stochastic dependence on the\ntime variable t:i.e., the state of a spin lying on the k-\nth lattice site was represented by a two-valued stochastic\nfunction σk(t)\nHI=−N/summationdisplay\nk=1/parenleftbig\nJIσkσk+1+gµBHe−iωtσk/parenrightbig\n, σk(t) =±1.\n(1)\nJIis the exchange coupling constant, that favors nearest\nneighboring spins to lie parallel ( JI>0, ferromagnetic\nexchange) or antiparallel ( JI<0, antiferromagnetic ex-\nchange); gis the Land´ e factor of each spin, and µBthe\nBohr magneton. In the original paper6a 1D lattice of\nequivalent and collinear spins was studied; there the re-\nsponse to a time-dependent magnetic field H( t), applied\nparallel to the axis of spin quantization and oscillating\nwith frequency ω, as in typical a.c.susceptibility exper-\niments, was also considered.\nIn order to investigate the phenomena of slow relax-\nation (for H=0) and resonant behavior of the a.c.sus-\nceptibility (for H ∝ne}ationslash= 0) in molecular SCM’s, we are going\nto generalize the Glauber model (1) accounting for non-\ncollinearity of local anisotropy axes and crystallographic\n(laboratory) frame. To this aim, we adopt the following\nmodel Hamiltonian\nH=−N/summationdisplay\nk=1/parenleftbig\nJIσkσk+1+GkµBHe−iωtσk/parenrightbig\n, σk(t) =±1.\n(2)\nJIis an effective Ising exchange coupling that can ap-\nproximately be related to the Hamiltonian parameters\nof a real SCM23,24: see later on the discussion in Sec-\ntion V. Like in the usual Ising-Glauber collinear model\n(1), the spins in Eq. (2) are described by classical, one-\ncomponent vectors that are allowed to take two integer\nvaluesσk(t) =±1, but now the magnetic moments may\nbe oriented along different directions, ˆzk, varying from\nsite to site. Within this scheme, the Land´ e tensor of a\nspin on the k-th lattice site has just a non-zero compo-\nnent,g/bardbl\nk, along the local easy anisotropy direction ˆzk.\nDenoting by ˆeHthe direction of the oscilla.ting magnetic\nfield,H(t)=He−iωtˆeH, we define the generalized Land´ e\nfactorGkappearing in Eq. (2) as\nGk=g/bardbl\nkˆzk·ˆeH (3)i.e., a scalar quantity that varies from site to site. Fol-\nlowing Glauber6, we define the single spin expectation\nvaluesk(t) =∝an}bracketle{tσk∝an}bracketri}htt, where brackets denote a proper en-\nsemble average, and the stochastic magnetization along\nthe direction of the applied field\n∝an}bracketle{tM∝an}bracketri}htt=µBN/summationdisplay\nk=1Gk∝an}bracketle{tσk∝an}bracketri}htt=µBN/summationdisplay\nk=1Gksk(t).(4)\nThe basic equation of motion of the Glauber model12,14\nreads\nd\ndtsk(t) =−2∝an}bracketle{tσkwσk→−σk∝an}bracketri}ht, (5)\nin which wσk→−σkrepresents the probability per unit\ntime to reverse the k-th spin, through the flip + σk→\n−σk. For a system of Ncoupled spins, this probability\nis affected by the interaction with the other spins, with\nthe thermal bath and, possibly, with an external mag-\nnetic field. Among all possible assumptions for the tran-\nsition probability wσk→−σkas a function of the N+ 1\nvariables11,19,25,26{σ1,...,σk,...,σN,t}, again following\nGlauber6we require wσk→−σkto be independent of time\nand to depend only on the configuration of the two near-\nest neighbors of the k-th spin. In zero field, these re-\nquirements are fulfilled by\nwH=0\nσk→−σk=1\n2α/bracketleftbigg\n1−1\n2γσk(σk−1+σk+1)/bracketrightbigg\n(6)\nwhile in the presence of an external field\nwH\nσk→−σk=wH=0\nσk→−σk(1−δkσk), (7)\nis usually chosen; the attempt frequency1\n2α(i.e., the\nprobability per unit time to reverse an isolated spin) re-\nmains an undetermined parameter of the model; γac-\ncounts for the effect of the nearest neighbors; the param-\netersδkhave the role of stabilizing the configuration in\nwhich the k-th spin is parallel to the field, and destabiliz-\ning the antiparallel configuration. Thanks to the partic-\nular choices (6) and (7) for the transition probability, by\nimposing the Detailed Balance conditions6it is possible\nto express γandδkas functions of the parameters in the\nspin Hamiltonian (2)\nγ= tanh(2 βJI), δk= tanh(βGkµBH).(8)\nwhereβ=1\nkBTis the inverse temperature in units of\nBoltzmann’s constant. Another advantage of Glauber’s\nchoices (6) and (7) is that the equation of motion (5)\ntakes a simple form. In particular, for H = 0, Eq. (5)\nwith the choice (6) becomes\nds(t)\ndt=−αAs(t), (9)\nwheres(t) denotes the vector of single-spin expectation\nvalues{s1(t), s2(t),···, sN(t)}andAis a square N×N3\nsymmetric matrix, whose non-zero elements are A k,k=1\nand A k,k−1=Ak,k+1=−γ\n2, with A 1,N=AN,1=−γ\n2if peri-\nodic boundary conditions are assumed for the N-spin\nchain. A closed solution of this set of first-order differen-\ntial equationscan be obtained expressingthe expectation\nvalue of each spin, sk(t), in terms of its spatial Fourier\nTransform (FT) /tildewidesq\nsk(t) =/summationdisplay\nq/tildewidesqeiqke−λqt. (10)\nSubstituting Eq. (10) into Eq. (9), one readily obtains\nthe dispersion relation\nλq=α(1−γcosq), q=2π\nNn (11)\nwithn= 0,1,..., N−127. For ferromagnetic cou-\npling (JI>0, hence γ >0) the smallest eigenvalue\nλq=0=α(1−γ) occurs for n=0, independently of the\nnumber of spins Nin the chain. For antiferromagnetic\ncoupling ( JI<0, hence γ <0) andNeven, the smallest\neigenvalue λq=π=α(1− |γ|) occurs for n=N\n2; while in\nthe case of Nodd, the smallest eigenvalue corresponds\ntoα/bracketleftbig\n1−|γ|cos/parenleftbigπ\nN/parenrightbig/bracketrightbig\n, thus depending on the number of\nspins in the antiferromagnetic chain28. The characteris-\ntic time scales of the system, τq, are given by\nτq=1\nλq=1\nα(1−γcosq). (12)\nAt finite temperatures the characteristic times τqare fi-\nnite because |γ|<1; forT→0 one has that 1 −|γ|van-\nishes irrespectively of the sign of JI, because γ→JI\n|JI|=\n±1. Thus, for H= 0, there is one diverging time scale\nin theT→0 limit:τq=0for ferromagnetic coupling and\nτq=πfor antiferromagneticcoupling (and even N). In the\npresence of a non-zero, oscillating field H( t) = He−iωt,\nthe equation of motion (5) with the choice (7) takes a\nform (see Eq. 24 in Section IV later on) which can still\nbe solved, though in an approximate way6, for a suffi-\nciently weak intensity of the applied magnetic field.\nIII. RELAXATION OF THE MAGNETIZATION\nIN ZERO FIELD\nThe originalGlauber model was formulated for a chain\nof collinear spins with the same Land´ e factors: i.e., in\nEq. (2) one has Gk=g,∀k= 1,···,N. Assuming\nthat the system has been fully magnetized by means of\na strong external field, one can study how the system\nevolves if the field is removed abruptly. This corresponds\nto take a fully saturated initial condition\nsk(0) = 1,∀k. (13)\nIn ferrimagnetic chains, on the other hand, a “par-\ntial” saturation can be reached, provided the antiferro-\nmagnetic coupling ( JI<0) between nearest neighbors is“strongenough”, in a sense that will be clarified later on.\nIn fact, if the Land´ efactorsfor odd and even sites arenot\nequal (go∝ne}ationslash=ge), through the application of an opportune\nfield the sample can be prepared in a configuration with\n/braceleftBigg\nsk(0) = +1 ,fork= 2r+1 (kodd)\nsk(0) =−1,fork= 2r(keven).(14)\nWith respect to the case considered by Glauber, it is\nconvenient to separate explicitly the expectation values\nof the odd sites, s2r+1(t), from those of the even sites,\ns2r(t). Thus, for H = 0, the set of Nequations of motion\n(9) can be rewritten as\n/braceleftBigg\nd\ndts2r=−α/bracketleftbig\ns2r+1\n2γ(s2r+1+s2r−1)/bracketrightbig\nd\ndts2r+1=−α/bracketleftbig\ns2r+1+1\n2γ(s2r+s2r−2)/bracketrightbig(15)\nIn the following, the solutions of (15) will be found using\ntwo different approaches that yield identical results.\nA. The Generating Function approach\nThe Generating Function approach, which closely fol-\nlows the original Glauber’s paper, is exposed in detail in\nAppendix. Here, in order to distinguish between the fer-\nromagnetic and ferrimagnetic relaxations, we specialize\nthe general solution, Eq. (A5) and Eq. (A6), to the two\ndifferentkindsofinitialconditions, Eq.(13)andEq.(14).\nIn both cases, we will assume that the exchange cou-\nplingJIis negative. The “partially saturated” configu-\nration, Eq. (14), reflects a typical experimental situation,\nin which the antiferromagnetic coupling is much bigger\n(JI≈100÷1000 K) than the Zeeman energy associated\nwith accessible magnetic fields. On the other hand, the\ninitial configurationwith allthe spins alignedin the same\ndirection, Eq. (13), clearly reflects the experimental situ-\nation of a fully saturated sample. This condition is easily\nobtained for ferromagnetic coupling ( JI>0), while it\nmay require very strong fields (eventually unaccessible)\nfor antiferromagnetic coupling ( JI<0).\nLet us startfrom the saturatedconfiguration, Eq.(13).\nSubstituting the initial condition sk(0) = 1 for all kin\nboth (A5) and (A6), we obtain\n\n\ns2r(t) =e−αt+∞/summationtext\nm=−∞/bracketleftbig\nI2(r−m)(γαt)+I2(r−m)−1(γαt)/bracketrightbig\ns2r+1(t) =e−αt+∞/summationtext\nm=−∞/bracketleftbig\nI2(r−m)(γαt)+I2(r−m)+1(γαt)/bracketrightbig\n.\nHence, exploiting the property (A4) of the Bessel func-\ntions (taking y= 1), and redefining the sums by a unique\nindexj, we get\n\n\ns2r(t) =e−αt+∞/summationtext\nj=−∞Ij(γαt) =e−α(1−γ)t\ns2r+1(t) =e−αt+∞/summationtext\nj=−∞Ij(γαt) =e−α(1−γ)t.(16)4\nThis meansthat, startingwith all the spinsalignedin the\nsamedirection,eachspinexpectationvalue(bothoneven\nand odd sites) decays obeying a mono-exponential law\nwith relaxation time τq=0= [α(1−γ)]−1, which is just\nthe characteristic time scale obtained as the inverse of\nthe dispersion relation λqwith zero wave number q= 0,\nseeEq. (11). Noticethat τq=0candivergefor T→0only\nin the case of ferromagnetic coupling, JI>0 (γ >0).\nLet us now consider the partially saturated configu-\nration, Eq. (14), in which sk(0) = 1 for kodd and\nsk(0) =−1 forkeven. Substituting these initial con-\nditions in both (A5) and (A6), we obtain\n\n\ns2r(t) =−e−αt+∞/summationtext\nm=−∞/bracketleftbig\nI2(r−m)(γαt)−I2(r−m)−1(γαt)/bracketrightbig\ns2r+1(t) =e−αt+∞/summationtext\nm=−∞/bracketleftbig\nI2(r−m)(γαt)−I2(r−m)+1(γαt)/bracketrightbig\nand, still exploiting the property (A4) (but now for y=\n−1), we get\n\n\ns2r(t) =−e−αt+∞/summationtext\nj=−∞(−1)jIj(γαt) =−e−α(1+γ)t\ns2r+1(t) =e−αt+∞/summationtext\nj=−∞(−1)jIj(γαt) =e−α(1+γ)t.\n(17)\nAlso in this case all the spins of the system relax with\na mono-exponential law, but now the relaxation time is\nτq=π= [α(1+γ)]−1, which corresponds to the inverse of\nthe eigenvalue λqwith wave number q=π, see Eq. (11).\nNotice that τq=πcan diverge for T→0 only in the case\nof antiferromagnetic coupling, JI<0 (γ <0).\nSummarizing, according to the sign of the exchange\nconstant, both time scales τq=0(forJI>0) andτq=π\n(forJI<0) diverge in the low temperature limit T→0,\nfollowing an Arrhenius law\nτ=1\n2αe4β|JI|(18)\nwith energy barrier 4 |JI|(slow relaxing mode). It is\nworth noting that the remaining relaxation times, given\nby the inverse of the eigenvalues in Eq. (11) with q∝ne}ationslash= 0\nandq∝ne}ationslash=π, always remain of the same order of mag-\nnitude as α−1(fast relaxing modes). This time scale is\ntypically very small ( ∼ps) in real systems, and negligible\nwith respect to the characteristic times involved in any\nexperimental measurement we refer to.\nB. The Fourier Transform approach\nThe solutions, (16) and (17), to the set of equations\n(15) can alternatively be deduced within the Fourier\nTransform (FT) formalism, which has already been ex-\nploited to obtain the dispersion relation (11). Recalling\nthe definition (10) of sk(t) and its spatial FT\n/tildewidesq=1\nN/summationdisplay\nksk(t)e−iqkeλqt, (19)we evaluate /tildewidesqat time t= 0,/tildewidesq=1\nN/summationtext\nksk(0)e−iqk,\nfor the two initial conditions of interest, (13) and (14).\nStarting from the all-spin-up configuration, Eq. (13), we\nhave\n/tildewidesq=1\nN/summationdisplay\nke−iqk=δq,0. (20)\nHence the solution for the expectation value of a spin\nlocalized on the klattice site at time tis\nsk(t) =/summationdisplay\nqδq,0eiqke−λqt=e−λq=0t,(21)\nwhich is identical to (16) since λq=0=α(1−γ).\nStarting from the partially saturated configuration,\nEq. (14), it is useful to rewrite it as sk(0) =−eiπk, so\nthat the FT at t= 0 is\n/tildewidesq=−1\nN/summationdisplay\nkeiπke−iqk=−δq,π.(22)\nHence the solution is readily obtained\nsk(t) =−/summationdisplay\nqδq,πeiqke−λqt=−eiπke−λq=πt,(23)\nwhich is identical to (17) since λq=π=α(1+γ).\nFinally, we observe that Eqs. (21) and (23) hold even\nfor a ring with a finitenumberNof spins, while Eqs.\n(16) and (17) were obtained in the infinite-chain limit.\nC. Slow versus fast relaxation of the spontaneous\nmagnetization\nThe expectation values, sk(t), of spins localized on\nthe even and odd sites of a linear lattice at time t,\ncomputed either with the Generating Functions or the\nFourier Transform approach, have been shown to dis-\nplay a mono-exponential relaxation, see Eqs. (16) and\n(17), with different time scales, τq=0= [α(1−γ)]−1and\nτq=π[α(1 +γ)]−1respectively, depending on the differ-\nent initial conditions, Eqs. (13) and (14). As a conse-\nquence, also the macroscopic magnetization, expressed\nby Eq. (4), displays the same mono-exponential relax-\nation as the single site quantities sk(t).\nA chain in which all the magnetic moments are equal\ncan be prepared only in the saturated initial configu-\nration, Eq. (13), with all the spin aligned in the same\ndirection, through the application of an external field.\nThus, when the field is abruptly removed, such a sys-\ntem will relax slowly at low temperature only if the ex-\nchange coupling is ferromagnetic ( JI>0). In contrast, if\nthe exchange coupling is antiferromagnetic ( JI<0) and\nthe chain is “forced” in the saturated state by a strong\napplied magnetic field, the system will relax very fast\n(in a typical time of the order of α−1) when the field\nis removed. Let us now discuss how these results, first\nobtained by Glauber6, are generalized to the case of a5\nchain in which the magnetic moments are collinear, but\nnotequal on each site.\nAs pointed out in the introduction, a model with an-\ntiferromagnetic coupling ( JI<0) but non-compensated\nmagnetic moments on the two sublattices is more akin\nto real SCM’s1,3,29. Yet it is very interesting since, de-\npending on the intensity of the applied magnetic field,\nthe system can be prepared either in the saturated ini-\ntial configuration, Eq. (13), where all spins are parallel\nto each other, or in the partially saturated one Eq. (14),\nwhere nearest neighbors are antiparallel. In the former\ncase, a very strong field is required in order to overcome\nthe antiferromagnetic coupling between nearest neigh-\nbors; once the field is removed, the relaxation of the\nmagnetizationis expected to be fast at lowtemperatures,\non the basis of the solution (16). In the latter case, the\npartiallysaturatedinitial configuration(14) can easilybe\nobtainedthroughtheapplicationofasmaller,experimen-\ntally accessible magnetic field; when the field is abruptly\nremoved, the relaxation is expected to be slow accord-\ning to the solution (17). The solution (17) justifies the\nobservation of SCM behavior in ferrimagnetic quasi-1D\ncompounds like CoPhOMe29(see Sect. V).\nSummarizing, we have found that when a collinear\nferrimagnetic chain is prepared in an initial state –\nfully or partially saturated depending on the intensity\nof the applied magnetic field – once the field is re-\nmoved abruptly, the spin system can show fast or slow\nrelaxation, respectively. Fast relaxation corresponds to\nstrongerfields; unfortunatelyforthe quasi-1Dchaincom-\npound CoPhOMe29,30, the antiferromagnetic exchange\nconstant is so large ( |JI| ∼100K) that the realization\nof a fully saturated initial configuration would require a\nvery high, almost unaccessible field ( ∼1000 kOe). Thus\nthis compound is not a good candidate for such a kind\nof experiments31.\nIV. MAGNETIC RESPONSE TO AN\nOSCILLATING MAGNETIC FIELD\nIn the presence of a magnetic field H, the transition\nprobability to be put in the equation of motion (5) is\nwH\nσk→−σk, defined in Eq. (7). One obtains\ndsk(t)\ndt=−α/braceleftBig\nsk(t)−γ\n2[sk+1(t)+sk−1(t)]\n+γδk\n2[∝an}bracketle{tσkσk+1∝an}bracketri}htt+∝an}bracketle{tσk−1σk∝an}bracketri}htt]−δk/bracerightBig\n(24)\nthat differs from Eq. (15), considered earlier for H= 0, in\nthe presence of both a non-homogeneous term, δk, and\nthe time-dependent pair-correlation functions ∝an}bracketle{tσkσk±1∝an}bracketri}htt.\nThe latter ones, assuming that the field is so weak to\ninduce just small departures from equilibrium, can be\napproximated by their time-independent counterparts32\n∝an}bracketle{tσkσk+1∝an}bracketri}htt=∝an}bracketle{tσk−1σk∝an}bracketri}htt≈tanh(βJI)≡η.(25)As it is usual in a.c.susceptibility measurements,\nwe also assume the time-dependent magnetic field\nH(t)=He−iωtˆeH, oscillating at frequency ω, to be weak\nso that the δkparameters can be linearized\nδk= tanh(βµBGkH(t))≈βµBGkH(t).(26)\nThe system of equations of motion (24) then takes the\nform\ndsk(t)\nd(αt)=−sk(t)+γ\n2[sk+1(t)+sk−1(t)]\n+βf(βJI)µBGkH(t), (27)\nwhere\nf(βJI) = 1−γη=1−η2\n1+η2(28)\nis a function of the reduced coupling constant βJIand\nwe have taken into account that γ= 2η/(1+η2). After\na brief transient period, the system will reach the sta-\ntionary condition in which the magnetic moment of each\nspin oscillates coherently with the forcing term at the\nfrequency ω. Expressing the expectation value of a spin\non thek-th lattice site, sk(t), through its spatial FT, /tildewidesq,\nthe trial solution is\nsk(t) =/summationdisplay\nq/tildewidesqeiqke−iωt. (29)\nSubstituting the latter in the system (27) we get\n/tildewidesq=βf(βJI)µBHα/tildewideGq\nα(1−γcosq)−iω,(30)\nwhere/tildewideGqis the FT of Gk:\n/tildewideGq=1\nNN/summationdisplay\nk=1e−iqkGk. (31)\nThe average of stochastic magnetization can readily be\nobtained from (4) as\n∝an}bracketle{tM∝an}bracketri}htt=µBe−iωtN/summationdisplay\nk=1/summationdisplay\nqq′/tildewideGq/tildewidesq′eiqkeiq′k,(32)\nwhich accounts for non-collinearity of local anisotropy\naxes with respect to the field direction. Performing the\nsum over all the lattice sites ( kindices) yields a factor\nNδq,−q′in Eq. (32); substituting the expression (30) for\n/tildewidesq, one obtains\n∝an}bracketle{tM∝an}bracketri}htt=Nµ2\nBβf(βJI)He−iωt/summationdisplay\nqq′α/tildewideGq/tildewideGq′δq,−q′\nα(1−γcosq)−iω.\n(33)\nThen, considering that /tildewideGq/tildewideG−q=|/tildewideGq|2, thea.c.suscep-\ntibility is finally obtained dividing (33) by H e−iωt\nχ(ω,T) =Nµ2\nBβf(βJI)/summationdisplay\nqα|/tildewideGq|2\nα(1−γcosq)−iω.(34)6\nFIG. 1: Temperature dependence of the imaginary part of the c omplex susceptibility, Eq. (38), of a collinear one-dimens ional\nIsing model with alternating spins. Resonant behavior in re sponse to an oscillating magnetic field is possible, at low fr equency,\nonly when magnetic moments are uncompensated (a,c,d), whil e a broad peak is found when the net magnetization is zero (b).\n(The curves refer to reduced frequency ω/α= 0.001)\nIn principle, the a.c.susceptibility of a chain with N\nspinsadmits Npoles, correspondingtothe Neigenvalues\nλqin Eq. (11). Each mode is related to a different time\nscaleτq= 1/λq. Inpractice, notallthetimescaleswillbe\ninvolved in the complex susceptibility χ(ω,T), butonly\nthe ones selected by /tildewideGq. A result similar, at first glance,\nto Eq. (34) was deduced by Suzuki and Kubo27, but in\ntheir case the relationship was between the time scale τq\nand the wave-vector-dependent susceptibility χ(q,ω). In\ncontrast, in an a.c.susceptibility experiment only the\nzero-wave-vector susceptibility χ(q= 0,ω) is accessible;\nthe peculiarity of Eq. (34) is that other time scales, dif-\nferent from τq=0, can be selected thanks to the depen-\ndence of the gyromagnetic factors Gkand of the local\nanisotropy axes on the site position k. This is the main\nresult of our study and will be clarified hereafter through\na few examples.\nA. The a.c.susceptibility of a collinear Ising\nferrimagnetic chain\nLet us start considering the case of a one-dimensional\nIsing model with two kinds of spins (aligned parallel on\nantiparallel to the chain axis) alternating on the odd\nand even magnetic sites of the lattice with Land´ e fac-torsG2r+1=goandG2r=ge(integerr), respectively.\nStrictly speaking, a collinear Ising ferrimagnet is char-\nacterized by an antiferromagnetic coupling JI<0, but\nalso the case JI>0 can be treated through Eq. (34).\nIn fact, since the local axis of anisotropy has the same\ndirection for all the spins, the FT of the site-dependent\nLand´ e factor is\n/tildewideGq=1\nNN/2/summationdisplay\nr=1[gee−iq2r+goe−iq(2r−1)]\n= (ge+eiqgo)1\nNN/2/summationdisplay\nr=1e−iq2r. (35)\nTaking into account that, in the presence of periodic\nboundary conditions, one has\nN/2/summationdisplay\nr=1e−iq2r=N\n2(δq,0+δq,π), (36)\nitfollowsthatthe onlynon-zerovaluesof /tildewideGqareforq= 0\nandq=π\n/tildewideG/bardbl\nq=1\n2[(ge+go)δq,0+(ge−go)δq,π].(37)\nThus, according to Eq. (34), the parallel a.c.suscepti-\nbility (∝bardbl=zz) of a collinear Ising chain with alternating7\nspins is\nχ/bardbl(ω,T) =Nµ2\nBβ f(βJI)\n×1\n4/bracketleftbigg(ge+go)2\n(1−γ)−i(ω\nα)+(ge−go)2\n(1+γ)−i(ω\nα)/bracketrightbigg\n(38)\nIt appears that both the relaxation times obtained by\nSuzuki and Kubo27for the ordinary and the staggered\nsusceptibility of the usual Ising model, namely τq=0=\n[α(1−γ)]−1andτq=π= [α(1 +γ)]−1respectively, do\ncoexist in the a.c.susceptibility (38). Notice that, in\ntheω→0 limit, the static susceptibility of the Ising\nferrimagnet in zero field17is recovered from Eq. (38),\nsince one hasf(βJI)\n1∓γ=1−η2\n1+η21\n1∓γ=e±2βJI.\nAs regards the dynamic response of the system to an\noscillating magnetic field applied along the chain axis,\ndepending on the sign of the effective exchange coupling\nconstant JI, the ferromagnetic ( ge+go) or the antifer-\nromagnetic ( ge−go) branch of the parallel susceptibility\n(38) are characterized by a diverging time scale at low\ntemperature. In particular, for a collinear Ising ferri-\nmagnet one has JI<0, so that τq=πis diverging, while\nτq=0is short (of the order of α−1, the attempt frequency\nof an isolated spin). Thus, for JI<0, a resonant be-\nhavior of the a.c.susceptibility versus temperature (at\nlow frequencies ω/α≪1) can only be observed in the\ncasege∝ne}ationslash=go(see Fig. 1d) when magnetic moments are\nuncompensated , while a broad peak is found in the case\nge=gowhen the net magnetization is zero (see Fig. 1b).\nClearly, for JI>0, a resonant peak is found in both\ncases (see Fig. 1a and 1c), because a net magnetization\nis always present in the system.\nSuch a resonant behavior of the a.c.susceptibility ver-\nsusT, in ferromagnetic33as well as in ferrimagnetic17\nIsing chains with single spin-flip Glauber dynamics, is a\nmanifestation of the stochastic resonancephenomenon34:\ni.e., the response of a set of coupled bistable systems to a\nperiodic drive is enhanced in the presence of a stochastic\nnoise when a matching occurs between the fluctuation-\ninduced switching rate of the system and the forcing fre-\nquency. In a magnetic chain, the role of stochastic noise\nis played by thermal fluctuations and the resonant peak\nin the temperature-dependence of the a.c.susceptibil-\nity occurs when the statistical time scale, associated to\nthe slow decay of the magnetization, matches with the\ndeterministic time scale of the applied magnetic field\nτq(Tpeak)≈1\nω. (39)\nB. The a.c.susceptibility of an n-fold helix\nNext, as an example of a non-collinear spin arrange-\nment, we consider a system of spins with the local axes\nof anisotropy arranged on an n-fold helix (see Fig. 2); θ\nis the angle that the local axes form with z, the unique\naxis of the helix ( i.e., the chain axis). In this case the\nLand´ e factors are equal on all lattice sites, but differentFIG. 2: Thick arrows denote the projections on the xyplane,\nperpendicular to the chain (helix) axis z, of magnetic mo-\nments in a one-dimensional Ising helimagnet, for different f old\nsymmetries ( n= 2,3,4,6). Dashed lines are the projections\nof the local axes of anisotropy, ˆzk.\nspins experience different fields because of the geometri-\ncal arrangement of magnetic moments. In the following\nwe will make the approximation that the Land´ e tensor\nof a spin on the k-th lattice site has just a non-zero com-\nponentgalong the easy anisotropy direction ˆzk, so that\nGk=gˆzk·ˆeH(see Eq. (3)). In the crystallographicframe\n(x,y,z), the directors ˆzkread (integer k)\nˆzk= sinθ/bracketleftbigg\ncos/parenleftbigg2πk\nn/parenrightbigg\nˆex+sin/parenleftbigg2πk\nn/parenrightbigg\nˆey/bracketrightbigg\n+cosθˆez.\n(40)\nLet us first consider the case of an oscillating magnetic\nfield H applied parallel to z, the helix axis. All the spins\nactually undergo the same field, and since Gk=gcosθ\nindependently of the lattice site k, the only peak in the\nFT/tildewideGqoccurs at q= 0\n/tildewideGz\nq=gcosθ δq,0(∀n). (41)\nFollowing the same procedure as in the previous para-\ngraph, the parallel a.c.susceptibility ( ∝bardbl=zz) takes an\nexpression (valid for any value of the fold index nof the\nhelix)\nχ/bardbl(ω,T) =Nµ2\nBβ f(βJI)g2cos2θ\n(1−γ)−i(ω\nα)(∀n) (42)\nthat differs from Glauber’s result for the collinear Ising\nchain6only by the geometrical factor cos2θ. For fer-\nromagnetic coupling, JI>0, the relaxation time τ0=\n[α(1−γ)]−1diverges as T→0, and a resonant behavior\nof thea.c.parallel susceptibility versus temperature is\nfound, at low frequency, when the oscillating field is ap-\nplied parallel to the helix axis, z, along which spins are\nuncompensated : see Fig. 3a, which refers to the case of a\ntwo-fold helix ( n= 2) .8\nFIG. 3: (color online) Temperature dependence of the imagin ary part of the parallel (42) and perpendicular (45) complex\nsusceptibility of an Ising chain with two-fold helical spin arrangement. The local axes ˆz1andˆz2were assumed to form an angle\nθ=π\n3withz, the chain axis (unique axis of the helix). Different curves r efer to different values of ω/α: 0.0001 (continuous, red\nline); 0.0002 (dashed, green line); 0.0005 (dashed single- dotted, blue line); 0.0010 (dashed double-dotted, violet l ine). Resonant\nbehavior in response to an oscillating magnetic field is poss ible, at low frequency, only for field applied in a direction w here\nmagnetic moments are uncompensated (a,c), while a broad pea k is found (b,d) when there is no net magnetization along the\nfield direction.\nLet us now consider the case of an oscillating magnetic\nfield H applied perpendicularly to the chain axis. In this\nconfiguration, it is useful to distinguish the case n= 2\nfrom the general case n >2.\n•n= 2\nIn this case, it is worth noticing that for H parallel\ntoy, one has identically Gr≡0 for any lattice\nsiter. Thus, /tildewideGy\nq≡0 and the corresponding a.c.\nsusceptibility is identically zero\nχyy(ω,T)≡0 (43)\n(not shown). In contrast, for H parallel to x, one\nhasGr=−gsinθon odd sites and Gr= +gsinθ\non even sites. The FT is\n/tildewideGx\nq=1\nNN/2/summationdisplay\nr=1gsinθ(−e−iq(2r−1)+e−iq2r)\n=gsinθ1\n2(δq,0+δq,π)(1−eiq)\n=gsinθ δq,π (44)\nwhere we have taken into account Eq. (36). Thus,\nfor ferromagnetic coupling, JI>0, the relaxationtimeτπ= [α(1+γ)]−1does not diverge as T→0,\nand the perpendicular a.c.susceptibility\nχxx(ω,T) =Nµ2\nBβf(βJI)g2sin2θ\n(1+γ)−i(ω\nα)(45)\ndoes not present a resonant behavior as a function\nof temperature; rather, it presents a broad max-\nimum (see Fig. 3b). Clearly, in the case of an-\ntiferromagnetic coupling, JI<0, the behavior of\nthe susceptibility components is reversed: a broad\nmaximum is found for the temperature dependence\noftheparallelsusceptibility χzz(ω,T)(seeFig. 3d),\nwhile a resonant behavior is found for the perpen-\ndicular susceptibility χxx(ω,T) (see Fig. 3c).\n•n >2\nIn this case, denoting by ˆex·ˆeHandˆey·ˆeHthe\ndirectors of the in-plane field, the FT’s of Gkare\ngiven by\n\n\n/tildewideGx\nq=1\nN(ˆex·ˆeH)gsinθ/summationtextN\nk=1cos/parenleftbig2πk\nn/parenrightbig\ne−iqk\n= (ˆex·ˆeH)gsinθ1\n2/parenleftBig\nδq,2π\nn+δq,−2π\nn/parenrightBig\n/tildewideGy\nq=1\nN(ˆey·ˆeH)gsinθ/summationtextN\nk=1sin/parenleftbig2πk\nn/parenrightbig\ne−iqk\n= (ˆey·ˆeH)gsinθ1\n2i/parenleftBig\nδq,2π\nn−δq,−2π\nn/parenrightBig.(46)9\nRemarkably, as just |/tildewideGq|2appears in Eq. (34), the\ngeneral result for the in-plane susceptibility turns\nout to be independent of the field direction. Thus,\nforn >2, the perpendicular ( ⊥)a.c.susceptibility\nof then-fold helix is given by\nχ⊥(ω,T) =Nµ2\nBβ f(βJI)\n×1\n2sin2θg2\n[1−γcos/parenleftbig2π\nn/parenrightbig\n]−i(ω\nα),(47)\nwhere we have exploited the fact that cos/parenleftbig\n−2π\nn/parenrightbig\n=\ncos/parenleftbig2π\nn/parenrightbig\nfor the term appearingat the denominator\nof Eq. (34).\nSummarizing, in the general case of an Ising chain\nwith an n-fold helical spin arrangement ( n≥2), we\nhave explicitly shown that a resonant behavior of the\na.c.susceptibility versus temperature, similar to the one\ndisplayed by ferromagnetic6,33and ferrimagnetic17Ising\nchains with collinear spins, is possible only for field ap-\nplied in a direction where magnetic moments are uncom-\npensated. In contrast, a broad peak is found when there\nis no net magnetization along the field direction.\nV. APPLICATION TO REAL SINGLE CHAIN\nMAGNETS\nIn this Section we will apply the developed formalism\nto some real compounds as representative realizations of\nSCM’s; for the three selected systems – we know this\nrestriction is far from being exhaustive1,3–a.c.suscep-\ntibility data on single crystal are available, which is a\nfundamentalrequirementforcheckingtheproposedselec-\ntion rules. The considered systems29,35,36are character-\nized by the alternation of two types of magnetic centers\nalong the chain axis, so that at least two spins per cell\nhave to be considered; moreover, the magnetic moments\nare not collinear, the dominant exchange interactions are\nantiferromagnetic and a strong single-ion anisotropy is\npresent, which favors magnetization alignment along cer-\ntain crystallographic directions ˆzk. The static properties\nof these compounds, like magnetization and static sus-\nceptibility, are generally well described using a classical\nHeisenberg model with an isotropic exchange coupling J\nand a single-ion anisotropy D. Thus, in order to describe\nthe dynamic behavior in response to a weak, oscillating\nmagnetic field by means of the previously developed the-\nory, it is necessary to relate the Hamiltonian parame-\nters of such a classical spin model to the exchange con-\nstantJIof the effective Ising model (2). In the following\nwe will show, through a few examples on real systems,\nthat indeed, depending on the geometry, selection rules\nare obeyed for the occurrence of slow relaxation of the\nmagnetization at low temperatures ( β|JI| ≫1), as well\nas for resonant behavior of the a.c.susceptibility as a\nfunction of temperature at low frequencies. As regards\nthe frequencies involved in an a.c.susceptibility exper-\niment on real SCM’s, generally1,3they lie in the range10−1÷104Hz, while the attempt frequency αis of the\norder of 1010÷1013Hz. Thus, for a typical experiment,\na resonant peak in the a.c.susceptibility can safely be\nobserved provided that at least one of the characteristic\ntime scales τqinvolved in (34) diverges at low T, in order\nfor the condition (39) to be satisfied.\nA. The MnIII-based Single Chain Magnet\nIn the one-dimensional molecular magnetic compound\nof formula [Mn(TPP)O 2PPhH]·H2O, obtained by re-\nacting Mn(III) acetate mesotetraphenylporphyrin with\nphenylphosphinic acid35, hereafter denoted by MnIII-\nbased SCM, the phenylphosphinate anion transmits a\nsizeable antiferromagnetic exchange interaction that,\ncombined with the easy axis magnetic anisotropy of the\nMnIIIsites, gives rise to a canted antiferromagnetic ar-\nrangement of the spins. The static single-crystal mag-\nnetic propertieswereanalyzedin the frameworkofaclas-\nsical spins Hamiltonian\nH=−N/2/summationdisplay\nr=1{JS2r−1·S2r+D/bracketleftbig\n(Sz1\n2r−1)2+(Sz2\n2r)2/bracketrightbig\n+e−iωtµBHαgαβ[Sβ\n2r−1+Sβ\n2r]} (48)\nwhereJ <0 is the antiferromagnetic nearest neighbor\nexchange interaction between S= 2 spins. D >0 is\nthe uniaxial anisotropy favoring two different local axes,\nalternating along odd and even sites respectively; both\naxes form an angle θ= 21.01owith the crystallographic\ncaxis, while they form opposite angles of modulus φ=\n56.55owith the aaxis (see Fig. 4). Thus we can write\nˆz2r−1= sinθcosφˆex−sinθsinφˆey+ cosθˆezandˆz2r=\nsinθcosφˆex+sinθsinφˆey+cosθˆez.\nA best fit of the static single-crystal magnetic suscep-\ntibilities, calculated via a Monte Carlo simulation35pro-\nvidesJ=−1.34 K and D= 4.7 K; the gyromagnetic\ntensorGαβis diagonal and isotropic with g/bardbl= 1.97.\nEquivalent results can be obtained calculating the static\npropertiesofmodel(48)viaatransfermatrixapproach37.\nSince the uniaxial anisotropy Dis rather strong with re-\nspect to the exchange coupling |J|, as a first approxima-\ntion one can assume the two sublattice magnetizations\nto be directed just along the two easy axes, ˆz2r−1and\nˆz2r, so that the chain system (48) can be described by a\nnon-collinear Ising model formally identical to Eq. (2),\nwith an effective38Ising exchange coupling JIand a gen-\neralized Land´ e factor Gkdefined as, respectively\nJI=JS(S+1)cos( ˆz2r−1·ˆz2r)\nGr=g/bardbl\nr/radicalbig\nS(S+1) (ˆzr·ˆeH). (49)\nDepending on the orientation of the oscillating mag-\nnetic field with respect to the crystallographic axes, the\nFT of the generalized Land´ e factor takes the following\nforms\n/tildewideGq=g/bardbl/radicalbig\nS(S+1)10\nFIG. 4: (color online) Disposition of local axes ( ˆz2r−1and\nˆz2r) and magnetic moments (red arrows) in the MnIII-based\nreal SCM, discussed in Sect. V.A, with antiferromagnetic\neffective Ising exchange coupling JI<0 . Right: Schematic\nview of the chain structure ( zis the chain axis) along the\ncrystallographic xaxis. Left: projections oflocal axes(dashed\nlines) and of magnetic moments (red arrows) in the xyplane,\nperpendicular to the chain axis.\n×1\nNN/2/summationdisplay\nr=1e−iq2r[eiq(ˆz2r−1·ˆeH)+(ˆz2r·ˆeH)]\n=g/bardbl/radicalbig\nS(S+1)\n\nsinθ1cosφ1δq,0,H∝bardblx\nsinθ1sinφ1δq,π,H∝bardbly\ncosθ1δq,0,H∝bardblz(chain axis)(50)\nThe corresponding a.c.susceptibility takes the expres-\nsion\nχ(ω,T) =Nµ2\nBβf(βJI)(g/bardbl)2[S(S+1)]\n×\n\nsin2θcos2φ1\n(1−γI)−i(ω\nα),H∝bardblx\nsin2θsin2φ1\n(1+γI)−i(ω\nα),H∝bardbly\ncos2θ1\n(1−γI)−i(ω\nα),H∝bardblz(chain axis)(51)\nTakingintoaccountthat, fortheMnIIISCMunderstudy,\nthe “true” exchange coupling, Jin Eq. (48), is antifer-\nromagnetic, and that the angle between the two easy\nanisotropy axes ˆz1andˆz2isδ= 34.6o<90o(see Fig.\n4, right), from Eq. (49) it follows that also the effective\nIsing exchange coupling is antiferromagnetic, JI<0. As\na consequence, in the low temperature limit β|JI| → ∞,\nthe relaxation time τq=πdiverges, while τq=0does not.\nThus, for low frequencies ω/α≪1, thea.c.susceptibil-\nity presents a resonant behavior only when the oscillat-\ning magnetic field is applied along the crystallographic yaxis,i.e.the direction, perpendicular to the chain axis,\nalong which the magnetizations of the two sublattices\nareuncompensated (see Fig. 4). In contrast, when H\nis applied parallel to z(the chain axis) or to x, namely\ntwo directions along which the magnetizations of the two\nsublattices are exactly compensated, no resonant behav-\nior is expected. These theoretical predictions turn out\nto be in excellent agreement with experimental a.c.sus-\nceptibility data35obtained in a single crystal sample of\n[Mn(TPP)O 2PPhH]·H2O, thus confirming that such a\nMnIII-based canted antiferromagnet is a bona fide SCM.\nB. The DyIII-based Single Chain Magnet\nThe molecular magnetic compound of formula\n[Dy(hfac) 3(NITPhOPh)], hereafter denoted by DyIII-\nbased SCM, belongs to a family of quasi one-dimensional\nmagnets in which rare earth ions (with spin S) and or-\nganic radical ions (with spin s= 1/2) alternate them-\nselvesalongthe chainaxis, z, which in this compound co-\nincides with the crystallographic baxis. Static measure-\nments in single crystal samples suggest36that there is an\nantiferromagnetic exchange interaction between neigh-\nboring DyIIIions, whose easy anisotropy axes are canted\nwithrespecttothechainaxisinsuchawaytogeneratean\nuncompensated moment along b, while the components\nin theacplane are compensated. Thus, as far as the\ndominant exchange interaction J <0 between DyIIIions\nis taken into account, the spin Hamiltonian of the sys-\ntem is quite similar to Eq.(48). However, with respect to\nthe MnIII-based chain, the crystal structure of the DyIII-\nbased SCM is more complicated, not only owing to the\npresence of two kinds of magnetic centers (the DyIIIions\nand the organic radical ions), but mainly because the\nsystem is formed by two different families of chains, with\ntwo almost orthogonal projections of the easy axes in\ntheacplane, perpendicular to the chain axis: this “ac-\ncidental” (in the sense that it is not imposed by symme-\ntry) orthogonality is the reason for the nearly isotropic\nmagnetic behavior displayed by the system within such\na plane36.\nWe adoptasimplified model formallyequivalentto Eq.\n(48). Taking into account only the dominant antiferro-\nmagnetic exchange interaction ( J <0) between neigh-\nboring DyIIIions (which indeed are next nearest neigh-\nbors in the real system) and their uniaxial anisotropy\n(D >0), the system can approximately be described by\nthe classical spins Hamiltonian (48), where now |Sk|= 1.\nBy means of a classical Transfer Matrix calculation, the\nstatic properties of the DyIII-based SCM turn out to be\nsatisfactorily fitted36byJ=−6 K,D= 40 K, g/bardbl= 10,\nwith the two easy anisotropy axes ˆz2r−1,ˆz2rforming\nequal angles θ≈75owith the chain axis z. (Notice that\nthe latter propertyholds true forboth families ofchains.)\nAlso in the case of the DyIII-based SCM, the uniaxial\nanisotropy Dturns out to be sufficiently strong with re-\nspect to the exchange coupling |J|in order to assume,11\nFIG. 5: (color online) Disposition of odd and even local axes\n(ˆz2r−1andˆz2r) and magnetic moments (thick arrows) in the\nDyIII-based real SCM, discussed in Sect. V.B, with ferro-\nmagnetic effective Ising exchange coupling JI>0 . Top:\nSchematic view of the chain structure ( zis the chain axis),\ndisplaying the two families of chains (A, with red magnetic\nmoments, and B, with green magnetic moments). Bottom:\nprojections of magnetic moments in the xyplane, perpendic-\nular to the chain axis.\nas a first approximation38, the two sublattice magneti-\nzations of DyIIIto be directed just along the two easy\naxes. Thus one can define an equivalent non-collinear\nIsing model (2), where the effective Ising exchange cou-\nplingJIand the generalized Land´ e factor Grare now\ndefined as\nJI=Jcos(ˆz2r−1·ˆz2r)\nGr=g/bardbl\nr(ˆzr·ˆeH). (52)\nDepending on the orientation of the oscillating magnetic\nfield with respect to the crystallographic axes, the FT of\nthe generalized Land´ e factor takes the form\n/tildewideGq=g/bardbl1\nNN/2/summationdisplay\nr=1e−iq2r[eiq(ˆz2r−1·ˆeH)+(ˆz2r·ˆeH)]\n∝g/bardbl/braceleftBigg\ncosθ δq,0,H∝bardblz(chain axis)\nsinθ δq,π,H⊥z.(53)\nIt is important to notice that this result holds true for\nbothfamilies(A,B)ofchains. Next, weobservethatsince\nin the DyIII-basedSCM, the spinsonopposite sublattices\narecoplanarwith the chain axis, the angle between ˆz2r−1\nandˆz2risjust2θ≈150o>90o. Takingintoaccountthat\nthe “true”exchangeconstantinEq. (48) isantiferromag-\nnetic,J <0, from Eq. (52) it follows that the effective\nIsingexchangecouplingisnowferromagnetic, JI>0(see\nFig. 5, top). As a consequence, in the low temperaturelimitβJI→ ∞, the relaxation time τq=0diverges, while\nτq=πdoes not. Thus, the a.c.susceptibility\nχ(ω,T)∝Nµ2\nBβf(βJI)(g/bardbl)2\n×/braceleftBigg\ncos2θ1\n(1−γI)−i(ω\nα),H∝bardblz(chain axis)\nsin2θ1\n(1+γI)−i(ω\nα),H⊥z(54)\nis expected to have a resonant behavior, for low frequen-\nciesω/α≪1, only when the oscillating magnetic field is\nappliedparalleltothechainaxis, z,alongwhichthemag-\nnetizations of the two sublattices are uncompensated (see\nFig. 5, top). Such a theoretical prediction turns out to\nbe in excellent agreementwith the experimental a.c.sus-\nceptibility data36obtained in a single crystal sample of\n[Dy(hfac) 3(NITPhOPh)] ∞, thus confirming that also the\nDyIII-based canted antiferromagnet is a bona fide SCM.\nTheonlyqualitativedifference, withrespecttotheMnIII-\nbased chain is that, due to the different geometry of the\nspin arrangement and of the local anisotropy axes with\nrespecttothechainaxis,theresonantbehaviorofthe a.c.\nsusceptibility is now observed for field applied parallel to\nthe chain axis, rather than perpendicular to it.\nC. The CoPhOMe (CoII-based) Single Chain\nMagnet\nIn the molecular magnetic compound of for-\nmula [Co(hfac) 2NITPhOMe], hereafter denoted by\nCoPhOMe29,30, the magnetic contribution is given by\nCobalt ions, with an Ising character and effective S=\n1/2, and by NITPhOMe organic radical ions, magnet-\nically isotropic and with s= 1/2. The spins are ar-\nranged on a helical structure, schematically depicted in\nFig. 6, right, whose projections in a plane perpendic-\nular to the helix axis z(coincident with the crystallo-\ngraphiccaxis), are represented in Fig. 6, left. The prim-\nitive magnetic cell is made up of three Cobalts (black\narrows) and three organic radicals (red arrows). Al-\nthough the effective spins of the two types of magnetic\ncenters have the same value, the gyromagnetic factors\nare different: gCo∝ne}ationslash=gR; thus, since the nearest neighbor\n(Cobalt-radical) exchange interaction is negative (and\nstrong,|J| ≈100 K)30, the sublattice magnetizations are\nnot compensated along z, whereas they are compensated\nwithin the xyplane perpendicular to the chain axis z.\nFor this compound, which was the first to display SCM\nbehavior29,30, static measurements on single-crystalsam-\nples has not been interpreted in terms of a simple model\nyet, due to the complexity of the system itself. Thus,\na relationship such as (49) and (52), which associate the\nIsing Hamiltonian (2) parameters( JIandGk) with those\nofa more realistic Hamiltonian, is still missing. However,\nthe dynamic behavior has been thoroughly investigated\ntreating - for the sake of simplicity - both the CoIIand\nthe organic radical spins as Ising variables, with σ=±1.12\nFIG. 6: (color online) Disposition of even and odd local axes\n(dashed lines) and magnetic moments (thick arrows) in the\nCoPhOMe real SCM, discussed in Sect. V.C, with antiferro-\nmagnetic effective Ising exchange coupling JI<0 . Right:\nSchematic view of the chain structure ( zis the chain axis)\nalong the crystallographic yaxis. Left: projections of local\naxes (dashed lines) and of magnetic moments (thick arrows)\nin thexyplane, perpendicular to the chain axis.\nThe effective Ising Hamiltonian reads\nH=−N/6/summationdisplay\nl=13/summationdisplay\nm=1{JIσl,2m[σl,2m−1+σl,2m+1]+e−iωtµBH\n[gRσl,2m−1(ˆz2m−1·ˆeH)+gCoσl,2m(ˆz2m·ˆeH)]}(55)\nwithlmagneticcellindexand msitelabelwithboundary\nconditions σl,7=σl+1,1. Since all the local axes ˆzk(k=\n1,···,6) form the same angle θ≈55owith the zaxis,\nwhen a magnetic field is applied along z, the FT of the\ngeneralized Land´ e factor is simply given by\n/tildewideG/bardbl\nq= cosθ1\nNN/2/summationdisplay\nr=1(gCoe−iq(2r−1)+gRe−iq2r)\n=cosθ\n2[(gCo+gR)δq,0+(gCo−gR)δq,π] (56)\nwhich, except for the prefactor cos θ, is quite similar to\nEq. (37) for the collinear Ising chain with alternating\nspins. Thus, the parallel a.c.susceptibility is\nχ/bardbl(ω,T) =Nµ2\nBβ f(βJI)cos2θ\n4\n×[(gCo+gR)2\n(1−γI)−i(ω\nα)+(gCo−gR)2\n(1+γI)−i(ω\nα)].(57)\nTaking into account that the effective exchange coupling\nof CoPhOMe is negative ( JI<0), the antiferromagnetic\nbranch of the parallel susceptibility is characterized by\na diverging time scale τq=π= [α(1+γI)]−1at low tem-\nperature, so that, for low frequencies ω/α≪1,χ/bardbl(ω,T)\ndisplays a resonant behavior.Let us now consider the case of a field applied in the\nplane perpendicular to the chain axis. For H ∝bardblx(see Fig.\n6, left) one has, letting k0=π\n3\nGx\n2r−1=gRsinθcos[k0(2r−1)]\nGx\n2r=gCosinθcos[k02r] (58)\nso that the FT takes the form\n/tildewideGx\nq= sinθ1\nNN/2/summationdisplay\nr=1e−iq2r/parenleftBig\ngCocos(k02r)\n+gReiq[cos(k0)cos(k02r)+sin(k0)sin(k02r)]/parenrightBig\n=1\n4sinθ[(gCo+gRei(q−k0))(δq,k0+δq,π+k0)\n+ (gCo+gRei(q+k0))(δq,−k0+δq,π−k0)] (59)\nwhere, as usual, we have exploited Eq. (36). Thus it\nfollows that\n/tildewideGq=±π\n3=sinθ\n4(gCo+gR)\n/tildewideGq=π±π\n3=sinθ\n4(gCo−gR).(60)\nThe corresponding relaxation times are τq=±π\n3=α\n1−1\n2γ\nandτq=±2π\n3=α\n1+1\n2γso that, summing the four contribu-\ntions we obtain the perpendicular a.c.susceptibility\nχ⊥(ω,T) =Nµ2\nBβ f(βJI)sin2θ\n8\n×/bracketleftbigg(gCo+gR)2\n(1−1\n2γ)−i(ω\nα)+(gCo−gR)2\n(1+1\n2γ)−i(ω\nα)/bracketrightbigg\n.(61)\nIn conclusion, for the six-fold helix model with alter-\nnating spins and Ising exchange coupling in Eq. (55),\nthe parallel and perpendicular components of the a.c.\nsusceptibility, χ/bardbl(ω,T) andχ⊥(ω,T), display a behavior\nsimilar to that of a ferrimagnetic chain with alternating\nspins (see (38)) and of an n-fold helical spin arrangement\nwith equivalent spins (see (47)), respectively. In spite of\nthe approximations involved in model (55) to describe\nthe real CoPhOMe molecular magnetic chain, the two\ncalculated susceptibilities (57) and (61), qualitatively re-\nproduce the dynamic behavior of this compound29,30. In\nfact, no out-of-phase a.c.susceptibility (imaginary part)\nis observed when the field is applied in the plane per-\npendicular to the chain axis, z, for the experimental fre-\nquencies (1 ÷105Hz)30. In contrast, when the oscillating\nfield is parallel to z, a resonant behavior is observed as\na function of temperature. Even though our theoretical\ntreatment holds only for small deviations from equilib-\nrium, it is worth mentioning that the absence of slow\nrelaxation for fields applied in the perpendicular plane is\nevidenced in the low temperature magnetization curve as\nwell: at low enough temperatures, a finite-area hysteresis\nloop is present only when a static field is applied parallel\nto the chain axis, while no hysteresis is observed in the\nin-plane magnetization curve29,30.13\nVI. CONCLUSIONS\nIn conclusion, in the framework of a one-dimensional\nIsing model with single spin-flip Glauber dynamics,\ntaking into account reciprocal non-collinearity of local\nanisotropy axes and the crystallographic (laboratory)\nframe, we have investigated: (i) the dynamics of mag-\nnetization reversal in zero field, and (ii) the response of\nthe system to a weak magnetic field, oscillating in time.\nWe have shown that SCM behavior is not only a feature\nof collinear ferro- and ferrimagnetic, but also of canted\nantiferromagnetic chains. In particular, we have found\nthat resonant behavior of the a.c.susceptibility versus\ntemperature in response to an oscillating magnetic field\nis possible, at low frequency, only for fields applied in a\ndirection where magnetic moments are uncompensated.\nIn contrast, a broad peak is expected when there is no\nnet magnetization along the field direction.\nThe role played by geometry in selecting the time\nscales involved in a process is an important and pecu-\nliarresult, typicalofmagneto-molecularapproachtolow-\ndimensional magnetism. In fact, magnetic centers with\nuniaxial anisotropyusually correspondto building blocks\nwith low symmetry39,40, which – in turn – often crystal-\nlize in more symmetric space groups, realizing a recip-\nrocal non-collinearity between local anisotropy axes as a\nnatural consequence1,2. Thus the family of real SCM’s,\nto which our model applies, does not restrict to ad-hoc\nsynthesized compounds but, instead, is expected to grow\nlarger in the future3. As a validity check of our selec-\ntion rules (as well as a tutorial exemplification), we have\nshown how our theory applies successfully to three differ-\nent molecular-based spin chains; when possible, we have\nput the parameters of our model Hamiltonian (2) in re-\nlationship with those of more general models, typically\nused tofit the staticpropertiesofthe correspondingcom-\npounds. Needless to say that the possibility of schema-\ntizing the chosen three compounds with Hamiltonian (2)\nrelies on the fact that at low enough temperatures they\nbehave as chains consisting of two-level units coupled by\na fully anisotropic exchange interaction. The latter as-\nsumption is expected to hold also for spin larger than\none-half in the presence of strong single ion anisotropy,\nprovided that domain walls still remain sharp24,38. In\nthis case each single magnetic center follows a thermally\nactivated dynamics, with an energy barrier ∆ 0, and well\nestablished heuristic arguments41suggest to replace the\nattempt frequency by α=α0e−β∆0.\nA naive application of our 3-fold-helix results (42)\nand (47) to the recently synthesized non-collinear Dy 3\ncluster would prevent the observation of slowrelaxation,\nwhile Single-Molecule-Magnet dynamics is indeed there\nobserved even in the presence of compensated magnetic\nmoments42. However, such a behavior in the classical\nregime,i.e.far from level crossings where underbarrier\nprocesses of quantum origin are important, is observed\nfor Dy 3in non-zero field and the resonant behavior is\ndue to a change of the relative population between thelowest and the first excited Kramers doublets of each Dy\nion: For sure this mechanism cannot be accounted for\nwhen dealing with two-level elementary variables, like σk\nin Hamiltonian (2). An extension of our model to mul-\ntivalued σk’s definitely deserve to be considered in the\nnext future.\nBeyondmolecularspinchains, ourapproachmightalso\nbe used to model monatomic nanowires showing slow re-\nlaxation ofthe magnetization at low temperatures43and,\npossibly, one-dimensional spin glasses44(provided that\nquenched disorder is somehow taken into account). In\nthis regard, the question of distinguishing between SCM\nand spin-glass behavior in quasi-1D systems is still a hot\ntopic of discussion45,46,47,48.\nAfter the successful organization of Single-Molecule\nMagnets onto surfaces49,50,51, the grafting of properly\nfunctionalized SCM’s on substrates represents a foresee-\nable goal as well as a fundamental step for their possi-\nble use as magnetic-memory units3. Technologies em-\nploying more traditional materials but based on alter-\nnative geometrical arrangement of magnetic anisotropy\naxes with respect to the switching field, such as in per-\npendicular recording52or processional switching53, are\nalready at the stage of forthcoming implementation in\ndevices. Were SCM’s to be considered as a possible route\nto tackle the main issues of high-density magnetic stor-\nage –i.e.optimization of the signal-noise ratio, thermal\nstability and writability52– the proposed selection rules\nfor slow relaxation, and related bistability, might find an\napplication in magnetic-memory manufacture as well.\nAcknowledgments\nWe wish to thank R. Sessoli and A. Rettori for stim-\nulating discussions, and J. Villain for interest and fruit-\nful suggestions in the early stages of this research work.\nFinancial support from ETHZ, the Swiss National Foun-\ndation, and Italian National Research Council is also ac-\nknowledged.\nAPPENDIX A: THE GENERATING FUNCTIONS\nAPPROACH\nAssuming periodic boundary conditions and defining\nthe two generating functions\nL(y,t) =+∞/summationdisplay\nr=−∞y2r+1s2r+1(t)\nG(y,t) =+∞/summationdisplay\nr=−∞y2rs2r(t), (A1)\nEqs.(15) taketheformoftwodifferentialequations(with\nthe dot indicating the first derivative with respect to the14\nadimensional variable αt)\n/braceleftBigg˙L(y,t) =−L(y,t)+1\n2γ(y+y−1)G(y,t)\n˙G(y,t) =−G(y,t)+1\n2γ(y+y−1)L(y,t).(A2)\nThe system (A2) can be decoupled through the substi-\ntution\nU(y,t) =L(y,t)+G(y,t)\nW(y,t) =L(y,t)−G(y,t), (A3)\nfrom which we directly get\n/braceleftBigg˙U(y,t) =−(1−ν)U(y,t)\n˙W(y,t) =−(1+ν)W(y,t),\nwithν=1\n2γ(y+y−1). The solutions of these two\nequations are U(y,t) =U(y,0)e−(1−ν)αtandW(y,t) =\nW(y,0)e−(1+ν)αtthat, exploiting the property\nexp/bracketleftbigg1\n2(y+y−1)x/bracketrightbigg\n=+∞/summationdisplay\nk=−∞ykIk(x) (A4)\nof the Bessel functions of imaginary argument Ik(x), can\nbe rewritten as\n\n\nU(y,t) =U(y,0)e−αt+∞/summationtext\nk=−∞ykIk(γαt)\nW(y,t) =W(y,0)e−αt+∞/summationtext\nk=−∞(−1)kykIk(γαt)\nPerforming the inverse transformation of (A3), the solu-\ntions to the system (A2) can be obtained\nL(y,t) =1\n2e−αt×\n+∞/summationdisplay\nk=−∞yk/bracketleftbig\nU(y,0)Ik(γαt)+(−1)kW(y,0)Ik(γαt)/bracketrightbig\nG(y,t) =1\n2e−αt×\n+∞/summationdisplay\nk=−∞yk/bracketleftbig\nU(y,0)Ik(γαt)−(−1)kW(y,0)Ik(γαt)/bracketrightbig\n.\nThen, separatingthe k-oddfromthe k-eventermsinboth\nsums, we get\nL(y,t) =e−αt+∞/summationdisplay\nr=−∞[y2rL(y,0)I2r(γαt)+y2r+1G(y,0)I2r+1(γαt)]\nG(y,t) =e−αt+∞/summationdisplay\nr=−∞[y2rG(y,0)I2r(γαt)\n+y2r−1L(y,0)I2r−1(γαt)].\nAccording to (A1), now L(y,0) andG(y,0) can be ex-\npressed again in terms of the initial single spin expecta-\ntion values s2r(0) ands2r+1(0) respectively\nL(y,t) =e−αt+∞/summationdisplay\nr=−∞\n×[y2r+∞/summationdisplay\nm=−∞y2m+1s2m+1(0)I2r(γαt)\n+y2r+1+∞/summationdisplay\nm=−∞y2ms2m(0)I2r+1(γαt)]\nPuttingk′=k+mwe have\nL(y,t) =e−αt+∞/summationdisplay\nk′=−∞y2k′+1+∞/summationdisplay\nm=−∞\n[s2m+1(0)I2(k′−m)(γαt)+s2m(0)I2(k′−m)+1(γαt)].\nComparing this latter result with the definition of L(y,t)\n(A1) and requiring for the terms corresponding to the\nsame power of yto be equal, an explicit function for the\nodd spin expectation values is readily obtained\ns2r+1(t) =e−αt+∞/summationdisplay\nm=−∞[s2m+1(0)I2(r−m)(γαt)\n+s2m(0)I2(r−m)+1(γαt)]. (A5)\nSubstituting L(y,0) eG(y,0) in the solution found for\nG(y,t) and performing similar passages, we obtain the\nexpectation value for even sites\ns2r(t) =e−αt+∞/summationdisplay\nm=−∞[s2m(0)I2(r−m)(γαt)\n+s2m+1(0)I2(r−m)−1(γαt)]. (A6)\n∗Electronic address: vindigni@phys.ethz.ch\n1C. Coulon, H.Miyasaka, and R.Cl´ erac, Struct. Bond. 122,\n163 (2006), and references therein.\n2H. MiyasakaandM. Yamashita, DaltonTrans., 399 (2007),\nand references therein.3L. Bogani et al., J. Mater. Chem. 18, 4750 (2008), and\nreferences therein.\n4H. B. Braun, Phys. Rev. Lett. 71, 3557 (1993).\n5H.B. Braun, J.Appl.Phys. 85, 6172(1999), andreferences\ntherein.15\n6R. J. Glauber, J. Math. Phys. 4, 294 (1963).\n7J. E. Anderson, J. Chem. Phys. 52, 2021 (1969).\n8S. Bozdemir, Phys. Status Solidi B 103, 459 (1981).\n9S. Bozdemir, Phys. Status Solidi B 104, 37 (1980).\n10J. L. Skinner, J. Chem. Phys. 79, 1955 (1983).\n11G. O.Berim andE.Ruckenstein, J. Chem.Phys. 119, 9640\n(2003).\n12J. J. Brey and A. Prados, Phys. Rev. E 53, 458 (1996).\n13R. Cordery, S. Sarker, and J. Toboshnik, Phys. Rev. B 24,\n5402 (1981).\n14J. Kamphorst Leal da Silva, A. G. Moreira, M. S. Soares,\nand F. C. S´ a Barreto, Phys. Rev. E 52, 4527 (1995).\n15R. B. Stinchcombe, J. E. Santos and M. D. Grynberg, J.\nPhys. A 31, 541 (1998).\n16M. Droz, J. K. L. da Silva and A. Malaspinas, Phys. Lett.\nA115, 448 (1986).\n17M. G. Pini and A. Rettori, Phys. Rev. B 76, 064407 (2007)\n[Erratum: Phys. Rev. B 76, e069903 (2007)].\n18E. Ising, Z. Phys. 31, 253 (1925).\n19M. Einax and M. Schulz, J. Chem. Phys. 115, 2282 (2001).\n20J. B. Goodenough, Magnetism and the Chemical Bond ,\nInterscience, New York, 1963.\n21J. B. Goodenough, J. Phys. Chem. Solids 6, 287 (1958).\n22J. Kanamori, J. Phys. Chem. Solids 10, 87 (1959).\n23L. Lecren et al., J. Am. Chem. Soc. 129, 5045 (2007).\n24A. Vindigni, Inorg. Chim. Acta 361, 3731 (2008).\n25B. U. Felderhof and M. Suzuki, Physica 56, 43 (1971).\n26J. C. Kimball, J. Stat. Phys. 21, 289 (1979).\n27M. Suzuki, R. Kubo, J. Phys. Soc. Japan 24, 51 (1968).\n28J. H. Luscombe, M. Luban, and J. P. Reynolds, Phys. Rev.\nE53, 5852 (1996).\n29A. Caneschi et al., Angew. Chem. Int. Ed. Ingl. 40, 1790\n(2001).\n30A. Caneschi et al., Europhys. Lett. 58, 771 (2002).\n31Itisworthnoticingthat, onceagoodcandidatewerefound,\nan experimental verification of the theoretical prediction\nmight be not so easy since the considered relaxation is\na strongly out-of-equilibrium process, while in our theory\nsmall departures from equilibrium were assumed.\n32K. Huang, Statistical Mechanics , J. Wiley and C., New\nYork, 1987.\n33J. J. Brey and A. Prados, Phys. Lett. A 216, 240 (1996).34L. Gammaitoni, P. H¨ anggi, P. Jung and F. Marchesoni,\nRev. Mod. Phys. 70, 223 (1998).\n35K. Bernot et al., J. Am. Chem. Soc. 130, 1619 (2008).\n36K. Bernot, Lanthanides in molecular magnetism: from\nmononuclear Single Molecule Magnets to Single Chain\nMagnets, Ph.D. thesis, INSA-Rennes, France (November\n2007).\n37R. Pandit and C. Tannous, Phys. Rev. B 28, 281 (1983).\n38For acollinear Heisenberg ferromagnet with exchange J\nand anisotropy D, the energy cost of a domain wall was\ncalculated and compared with the energies of a sharp wall\nand of a soliton, and a the crossover between the “sharp\nwall” regime ( J≪D) and the “broad wall” regime ( J≫\nD) was found to occur24forJ/D= 1.8. In principle, a\nsimilar calculation should be performed also for the non-\ncollinear model (48), in order to find the limits of validity\nfor the approximation made in Eq. (49).\n39D. Gatteschi et al., Science 265, 1054 (1994).\n40D. Gatteschi and R. Sessoli, Angew. Chem. Int. Ed. 42,\n268 (2003), and references therein.\n41C. Coulon, R. Cl´ erac1, L. Lecren, W. Wernsdorfer, and H.\nMiyasaka, Phys. Rev. B 69, 132408 (2004).\n42J. Luzon et al., Phys. Rev. Lett. 100, 247205 (2008).\n43P. Gambardella et al., Nature 416, 301 (2002).\n44J. A. Mydosh, Spin Glasses: An Experimental Introduc-\ntion, Taylor and Francis Ltd., London, 1993.\n45A. Maignan et al., Eur. Phys. J. B 15, 657 (2000).\n46S. J. Etzkorn, W. Hibbs, J. S. Miller, and A. J. Epstein,\nPhys. Rev. B 70, 134419 (2004).\n47L. Bogani, Magnetic and magneto-optical properties of\nmolecular compounds , Ph.D.thesis, DipartimentodiChim-\nica, Universit` a di Firenze, Italy (December 2005).\n48M.A. Girtu et al., J. Appl. Phys., 81, 4410 (1997).\n49A. Cornia et al., Angew. Chem. Int. Ed. Ingl., 42, 1645\n(2003).\n50A. N. Abdi et al., J. Appl. Phys. 95, 7345 (2004).\n51M. Mannini et al., Chem. Eur. J. 14, 7530 (2008).\n52S. N. Piramanayagam and K. Srinivasan, J. Mag. Mag.\nMat.,in press (2008).\n53C. H. Back et al., Science 285, 864 (1999)." }, { "title": "1304.2826v1.Magnetization_and_spin_dynamics_of_the_spin_S_1_2_hourglass_nanomagnet_Cu5_OH_2_NIPA_4_10H2O.pdf", "content": "Magnetization and spin dynamics of the spin S=1\n2hourglass nanomagnet\nCu5(OH) 2(NIPA) 4\u000110H 2O\nR. Nath,1A. A. Tsirlin,2, 3,\u0003P. Khuntia,2O. Janson,2, 3T. F orster,2, 4M. Padmanabhan,5\nJ. Li,6Yu. Skourski,4M. Baenitz,2H. Rosner,2and I. Rousochatzakis7,y\n1School of Physics, Indian Institute of Science Education and Research, Thiruvananthapuram-695016, India\n2Max Planck Institut f ur Chemische Physik fester Sto\u000be, N othnitzer Str. 40, 01187 Dresden, Germany\n3National Institute of Chemical Physics and Biophysics, 12618 Tallinn, Estonia\n4Dresden High Magnetic Field Laboratory, Helmholtz-Zentrum Dresden-Rossendorf, 01314 Dresden, Germany\n5School of Chemistry, Indian Institute of Science Education and Research, Thiruvananthapuram-695016, India\n6Department of Chemistry &Chemical Biology, Rutgers University, Piscataway, NJ 08854, USA\n7Institute for Theoretical Solid State Physics, IFW Dresden, 01171 Dresden, Germany\nWe report a combined experimental and theoretical study of the spin S=1\n2nanomagnet\nCu5(OH) 2(NIPA) 4\u000110H 2O (Cu 5-NIPA). Using thermodynamic, electron spin resonance and1H nu-\nclear magnetic resonance measurements on one hand, and ab initio density-functional band-structure\ncalculations, exact diagonalizations and a strong coupling theory on the other, we derive a micro-\nscopic magnetic model of Cu 5-NIPA and characterize the spin dynamics of this system. The ele-\nmentary \fve-fold Cu2+unit features an hourglass structure of two corner-sharing scalene triangles\nrelated by inversion symmetry. Our microscopic Heisenberg model comprises one ferromagnetic and\ntwo antiferromagnetic exchange couplings in each triangle, stabilizing a single spin S=1\n2doublet\nground state (GS), with an exactly vanishing zero-\feld splitting (by Kramer's theorem), and a very\nlarge excitation gap of \u0001 '68 K. Thus, Cu 5-NIPA is a good candidate for achieving long electronic\nspin relaxation ( T1) and coherence ( T2) times at low temperatures, in analogy to other nanomag-\nnets with low-spin GS's. Of particular interest is the strongly inhomogeneous distribution of the\nGS magnetic moment over the \fve Cu2+spins. This is a purely quantum-mechanical e\u000bect since,\ndespite the non-frustrated nature of the magnetic couplings, the GS is far from the classical collinear\nferrimagnetic con\fguration. Finally, Cu 5-NIPA is a rare example of a S=1\n2nanomagnet showing\nan enhancement in the nuclear spin-lattice relaxation rate 1 =T1at intermediate temperatures.\nPACS numbers: 75.50.Xx, 75.10.Jm, 75.30.Et, 76.60.Jx\nI. INTRODUCTION\nThe \feld of molecular nanomagnets has enjoyed an\nenormous experimental and theoretical activity over\nthe last few decades.1Owing to the nanoscopic size\nof their elementary magnetic units, these compounds\nprovide experimental access to a plethora of quantum\nmechanical (QM) e\u000bects, including quantum tunneling\nof the magnetization2,3or the N\u0013 eel vector,4quantum\nphase interference,5level-crossings and magnetization\nplateaux.6They also allow to probe on the macroscopic\nscale the crossover from quantum to classical physics.7\nFinally, molecular magnets are promising materials for\nspintronic applications8and quantum computing.9\nHere we report on the magnetic behavior, mi-\ncroscopic magnetic model, and spin dynamics of\nCu5(OH) 2(NIPA) 4\u000110H 2O, hereinafter referred to as\nCu5-NIPA, where the acronym NIPA stands for the\n5-nitro-isophtalic acid ligand. It is an hourglass-shaped\nmolecular magnet comprising \fve Cu2+spin-1\n2ions. The\nground state (GS) of this magnet has a low spin value\nS=1\n2with a very large spin gap of \u0001 '68 K. Therefore,\nCu5-NIPA behaves as a rigid spin S=1\n2entity in a wide\ntemperature range, and resembles other spin-1\n2molecu-\nlar magnets, such as V 6.10,11Given that both compounds\ncomprises=1\n2spins, we expect similarly long electron\nspin-phonon relaxation times T1, which allow for the ob-servation of rich hysteresis e\u000bects in pulsed \felds.10,12\nHowever, in contrast to V 6, here we do not expect abrupt\nsteps in the magnetization curve,10because the present\ncompound features an odd number of half-integer spins,\nthus the zero-\feld splitting vanishes exactly by Kramer's\ntheorem.\nIn analogy with other molecular magnets with low-\nspin GS's, such as iron-sulfur clusters,13{15heterometallic\nrings,16iron trimers,17V15,18and single-molecule mag-\nnets (SMM),19we expect that Cu 5-NIPA manifests also\nlong coherence times T2, which is a crucial step towards\nimplementations in quantum computing.20\nAnother attractive feature of Cu 5-NIPA is the pres-\nence of two corner-sharing scalene triangles, related\nby inversion symmetry.21,22The spin triangle is the\nmost elementary unit for highly frustrated magnetism,23\nwhile its chirality may induce a \fnite magnetoelec-\ntric coupling,24,25and may also be used as a qubit.26\nIn molecular magnetism, the spin triangle is com-\nmon in many compounds,1such as V 15,12,27Cu3\nclusters,28the chiral Dy 3cluster,29the cuboctahedron\nCu12La8,30as well as the giant icosidodecahedral ke-\nplerates Mo 72Fe30,31W72Fe30,32Mo72Cr30,33Mo72V3034\nand W 72V30,35which host highly frustrating kagome-like\nphysics.36\nThe distinct feature of the present compound is the\nstrong distortion of its regular spin triangles, featuringarXiv:1304.2826v1 [cond-mat.mtrl-sci] 10 Apr 20132\nb\nCu3\n217 K\n217 KCu1\nCu1\n62□K62□K\n/c4562 K\n/c4562 KCu2\nCu2a\nc\n32\n21\n1\nCC\nOO\nO/c109\n3-OH/c1093-OH\nO\nFIG. 1. (Color online) Left panel: crystal structure of Cu 5-NIPA showing the stacking of the Cu 5units (one unit is highlighted\nby gray shading) through the NIPA ligand molecules, into a rigid three-dimensional metal-organic framework, with the shortest\ndistance of 8.6 \u0017A between the magnetic units. Middle and right panels: structure of the Cu 5magnetic unit and relevant\nmagnetic interactions according to the \ft of the experimental susceptibility data (Fig. 13). The numbers 1 \u00003 denote the\ncrystallographic positions Cu1{Cu3, with Cu3 residing at the inversion center. Arrows show the classical ferrimagnetic ground\nstate.\nthree nonequivalent Cu positions, denoted by Cu1, Cu2,\nand Cu3 (see Fig. 1). Each spin triangle is centered by the\n\u00163-OH group37providing Cu{O{Cu superexchange path-\nways, while the carboxyl (COO\u0000) group of the NIPA lig-\nand creates an additional Cu1{Cu3 pathway (see Fig. 1,\nmiddle). This topology leads to three drastically di\u000berent\ninteractions, one of which is ferromagnetic (FM). Despite\nthe fact that these couplings do not compete with each\nother classically, the GS has strong QM character, which\nis partly re\rected in a strongly inhomogeneous distribu-\ntion of the magnetic moment over the \fve Cu2+spins.\nOf particular interest is our experimental \fnding of\nan enhancement of the1H nuclear spin-lattice relaxation\nrate 1=T1at a characteristic temperature slightly below\nthe spin gap ( T'40 K). While such an enhancement\nhas been repeatedly found in numerous antiferromag-\nnetic (AFM) homometallic38and heterometallic39rings\nof spinss >1\n2, it is very rare for s=1\n2. We argue that\nthe origin of the peak is the same in both cases, namely\nthe slowing down of the phonon-driven magnetization\n\ructuations.38,40{42However, there are several qualita-\ntive di\u000berences with the case of homometallic rings, re-\nlated to the sparse excitation spectrum of Cu 5-NIPA and\nthe presence of inequivalent Cu2+sites.\nThe organization of this article is the following. We\nbegin in Sec. II with details on the sample prepara-\ntion and methods. Next, we discuss bulk speci\fc heat\nand magnetization (Sec. III A), as well as local electron\nspin resonance (Sec. III B) and1H nuclear magnetic res-\nonance data (Sec. III C). The microscopic description of\nCu5-NIPA in terms of the isotropic Heisenberg model\nis based on ab initio density-functional band-structure\ncalculations (Sec. IV A) and facilitates the evaluation of\nmagnetic properties (local magnetizations, nature of the\nGS, etc.) using exact diagonalizations (Sec. IV B) and a\nstrong coupling theory (Sec. IV C). Finally, we conclude\nwith a discussion and some interesting perspectives of\nthis study in Sec. V.II. METHODS\nA powder sample of Cu 5-NIPA was prepared according\nto the procedure described in Ref. 21. Sample purity was\nchecked by powder x-ray di\u000braction (Huber G670 Guinier\nCamera, CuK \u000b1radiation, 2 \u0012= 3\u0000100\u000eangular range).\nThe magnetic susceptibility ( \u001f) was measured in the\ntemperature range 1 :8 K\u0014T\u0014400 K in an applied \feld\nof\u00160H= 1 T using the commercial Quantum Design\nMPMS SQUID. The magnetization isotherm Mvs.H\nwas measured at T= 2 K in \felds up to 14 T using the\nvibrating sample magnetometer (VSM) option of Quan-\ntum Design PPMS. Additionally, pulsed-\feld measure-\nments in \felds up to 60 T were performed at 1.4 K in the\nDresden High Magnetic Field Laboratory. Details of the\nmeasurement procedure are described in Ref. 43.\nThe heat capacity Cp(T) was measured on a\nsmall piece of pellet over the temperature range\n2 K\u0014T\u0014300 K in zero \feld using the Quantum De-\nsign PPMS.\nThe electron spin resonance (ESR) measurements were\nperformed at Q-band frequencies ( f= 34 GHz) using a\nstandard spectrometer together with a He-\row cryostat\nthat allows us to vary the temperature from 1.6 to 300 K.\nESR probes the absorbed power Pof a transversal mag-\nnetic microwave \feld as a function of a static and exter-\nnal magnetic \feld B. To improve the signal-to-noise ra-\ntio, we used a lock-in technique by modulating the static\n\feld, which yields the derivative of the resonance signal\ndP=dB .\nThe nuclear magnetic resonance (NMR) measurements\nwere carried out using pulsed NMR technique on1H nu-\nclei (nuclear spin I=1\n2and gyromagnetic ratio \rn=2\u0019=\n42:576 MHz/T) in the 2 \u0000230 K temperature range. The\nNMR measurements were performed at two di\u000berent ra-\ndio frequencies of 70 MHz and 38.5 MHz, which corre-\nspond to an applied \feld of about 1 :6608 T and 0 :9135 T,3\n600\n0.6\n0.4\n0.2\n0.00.81.0 dehydrated\npristine1.21.4\n/c99*T(emu□K/mol□Cu) 400\n200\n0\n0 100 100 10 2 200 300 400\nTemperature□(K) Temperature□(K)80010001200\ndehydrated□Cu -NIPA\n=□1.75 , =□9□K5\neff B/c109 /c113 /c109hydrated□Cu -NIPA5\neff B/c109 /c109 /c113=□1.75 , =□34□K\nground□state\n1/ (emu/mol□Cu)/c99/c451\n1\n2\n/c109 /c109 /c113eff B=□0.9 , =□0\nground□statehigh- paramagnet T\n1\n2\nFIG. 2. (Color online) Left panel: Inverse magnetic susceptibility of Cu 5-NIPA measured upon heating from 2 K (circles,\nhydrated form) and upon cooling from 400 K (triangles, dehydrated form). Lines show Curie-Weiss \fts to Eq. (1), as described\nin the text. Right panel: \u001f\u0003Tplot showing the formation of the\f\f1\n2istate in the pristine sample ( \u001f\u0003T=1\n4) and the high-\ntemperature paramagnetic regime of the dehydrated sample ( \u001f\u0003T=5\n4). Here,\u001f\u0003=\u001f\u0010\nNg2\u00162\nB\nkB\u0011\u00001\n,N=NA=5, andg= 2:2\naccording to ESR (Sec. III B). For the \u001f(T) plot, see Fig. 13.\nrespectively. Spectra were obtained by Fourier transform\nof the NMR echo signal. The1H spin-lattice relaxation\nrate, 1=T1, was measured by using the conventional sat-\nuration pulse sequence.\nThe electronic structure of Cu 5-NIPA was calculated in\nthe framework of density functional theory (DFT) using\ntheVASP code44for crystal structure optimization and\nFPLO45for the evaluation of magnetic parameters. The\ngeneralized gradient approximation (GGA)46exchange-\ncorrelation potential was augmented with the mean-\feld\nDFT+Ucorrection for Coulomb correlations in the Cu\n3dshell. The DFT+ Uparameters were chosen as Ud=\n9:5 eV (on-site Coulomb repulsion), Jd= 1 eV (on-site\nHund's exchange), and fully-localized-limit (FLL) \ra-\nvor of the double-counting correction, following earlier\nstudies of Cu2+-based magnets.47Reference calculations\nwith other exchange-correlation potentials and double-\ncounting corrections arrived at qualitatively similar re-\nsults. The reciprocal space was sampled by a kmesh\nwith 64 points in the \frst Brillouin zone. The conver-\ngence with respect to the kmesh was carefully checked.\nDetails of the computational procedure are reported in\nSec. IV A.\nThermodynamic and GS properties of the Cu 5\nmolecule were evaluated by full numerical diagonaliza-\ntions.\nIII. EXPERIMENTAL RESULTS\nA. Thermodynamic properties\n1. Magnetization\nThe magnetic susceptibility of Cu 5-NIPA was mea-\nsured for the as-prepared sample under \feld-cooling (FC)\ncondition upon heating from 1.8 K to 400 K. At 400 K,the sample was kept inside the SQUID magnetometer for\nabout 20 minutes, and \u001f(T) was measured again upon\ncooling from 400 K to 1.8 K. The drastic di\u000berence be-\ntween the heating and cooling curves (Fig. 2) indicates\nthe decomposition of Cu 5-NIPA shortly above room tem-\nperature. Indeed, the sample color changed from blue to\ngreen, and the weight loss of 13 % was detected. This\nweight loss is in good agreement with the preceding ther-\nmogravimetric data of Ref. 22 that reported the weight\nloss of 13.6 % at 380 \u0000400 K. Although the decomposi-\ntion process is tentatively ascribed to the release of water\nmolecules (expected weight loss of 13.15 %)22, no detailed\ninformation on the decomposed sample is available in the\nliterature.\nThe pristine sample shows a sharp decrease in the mag-\nnetic susceptibility (increase in 1 =\u001f) upon heating and a\nbend around 30 K followed by the nearly linear regime of\n1=\u001fabove 200\u0000250 K. In contrast, the decomposed sam-\nple remains paramagnetic down to at least 10 K. Below\n320 K, where the decomposition process is \fnished, the\ninverse susceptibility of dehydrated Cu 5-NIPA follows a\nstraight line that can be \ftted with the Curie-Weiss law\nin the 20\u0000320 K temperature range,48\n\u001f=C\nT+\u0012: (1)\nThe resulting Curie constant C'0:383 emu K/(mol Cu)\nleads to an e\u000bective moment \u0016e\u000b'1:75\u0016B, which is\nclose to 1.73 \u0016Bexpected for spin-1\n2. The Weiss temper-\nature is\u0012'9 K.\nThe inverse susceptibility of the pristine Cu 5-NIPA\nsample does not have a well-de\fned paramagnetic re-\ngion, because the data above 320 \u0000350 K are a\u000bected\nby the decomposition. A tentative Curie-Weiss \ft in\nthe 220\u0000320 K range yields an e\u000bective moment of\n\u0016e\u000b'1:75\u0016B(C= 0:383 emu K/(mol Cu)), which is\nsame as in the dehydrated sample. However, the Weiss4\n0.8\n0.4\n0.0\n5\nFiel d□(T)0 10Experiment\nBrillouin□function\n( =□2.38)g\nT=□1.8□K\n151.2\nMs=□1.19 /f.u. /c109B\nMagnetization□( /f.u.)/c109B\nFIG. 3. (Color online) Magnetization curve of Cu 5-NIPA\nmeasured at 1.8 K up to 14 T. The dashed line denotes the\nsaturation of the\f\f1\n2istate atMs= 1:19\u0016B/f.u. The solid\nline is the Brillouin function calculated at 1.8 K with g= 2:38.\ntemperature of \u0012'34 K in the pristine sample is notably\nlarger than \u0012'9 K observed in the dehydrated sample.\nThe conspicuous di\u000berence between the hydrated and\ndehydrated samples suggests a huge e\u000bect of dehydra-\ntion on the magnetism of the Cu 5molecule. The dehy-\ndration proceeds in a single step and results in the loss\nof all 10 water molecules per formula unit. As 4 out\nof these 10 molecules enter the \frst coordination sphere\nof Cu (Fig. 1, left), a large e\u000bect on the local environ-\nment of Cu sites and, thus, on the magnetism should\nbe expected.49Unfortunately, no structural data for the\ndehydrated version of Cu 5-NIPA are presently available.\nTherefore, we restrict ourselves to a detailed study of the\nhydrated compound.\nAt low temperatures, the inverse susceptibility of Cu 5-\nNIPA approaches another linear regime, with the Curie\nconstantC= 0:101 emu K/(mol Cu) ( \u0016e\u000b'0:90\u0016B)\nand vanishingly small \u0012'0:5 K. This low-temperature\nparamagnetic state can be also seen in the \u001f\u0003Tplot\n(\u001f\u0003=\u001f\u0000\nNg2\u00162\nB=kB\u0001\u00001is the reduced susceptibility),\nwhere\u001f\u0003Tapproaches the value of1\n4expected for the\ntotal spinS=1\n2per molecule (Fig. 2, right). Likewise,\nthe magnetization curve of Cu 5-NIPA (Fig. 3) saturates\natMs= 1:19\u0016B/molecule, which is, however, higher\nthan the free-electron value of 1 \u0016B/f.u. This di\u000berence\ncan be well accounted for by the large g-value of 2.38 ac-\ncording toMs=gS\u0016B. The same g-value explains the\nlow-temperature e\u000bective moment: C=Ng2\u00162\nB=3kB=\n0:106 emu K/(mol Cu), where we use N=NA=5 for\nthe magnetic moment of1\n2per Cu 5molecule. As we ex-\nplain later, this change in the g-value (Sec. IV B) can be\ntraced back to the presence of three non-equivalent Cu\npositions with di\u000berent g-tensor anisotropies. At room\ntemperature, the Cu spins are nearly independent, and\ntheirg-values average with same weights for all Cu posi-\ntions. At low temperatures, the powder averaging of the\ng-values is determined by the distribution of the magne-\ntization in the S=1\n2GS.\n4\n2\n0\n10 100 26\nC Tp/ (J□mol K )/c451/c452\nTemperature□(K)100 200 00400800\nCp(J□mol K )/c451/c45/c492 4 60.51.0\nC Tp/FIG. 4. (Color online) Speci\fc heat of Cu 5-NIPA divided\nby temperature ( Cp=T). The inset in the upper left corner\nmagni\fes the data at low temperatures. The inset in the\nbottom right corner shows the speci\fc heat ( Cp).\nOur results suggest that at low temperatures Cu 5-\nNIPA is in the paramagnetic state with the total moment\nofS=1\n2per Cu 5molecule. This state is further denoted\nas\f\f1\n2i. The experimental magnetization curve follows\nthe Brillouin function\nB(H) =g\u0016BS\u0002th\u0012g\u0016BSH\nkBT\u0013\n(2)\nwithS=1\n2,g= 2:38, andT= 1:8 K (Fig. 3). A marginal\ndeparture of the experimental curve toward higher \felds\nmay be due to anisotropies and/or intermolecular cou-\nplings. The maximum deviation of about 0.3 T puts an\nupper limit of about 0.5 K on the total energy of such\ncouplings. This low energy scale is in agreement with the\nlarge spatial separation between the molecules (Fig. 1,\nright).\nBelow 4 K, the susceptibility of Cu 5-NIPA departs\nfrom\u001f\u0003T=1\n4(Fig. 2, right). This decrease in \u001f\u0003Tmay\nindicate an evolution of the system toward a long-range\nmagnetic order between the Cu 5molecules. However, we\nwere unable to see any clear signatures of a magnetic\ntransition in the susceptibility data measured down to\n2 K.\n2. Speci\fc heat\nThe speci\fc heat ( Cp) of Cu 5-NIPA is smooth down\nto 2 K (Fig. 4). It steeply increases up to 60 \u000070 K\nand shows a slower increase at higher temperatures, as\nshown by a broad maximum in the temperature depen-\ndence ofCp=T. The signal is dominated by the phonon\ncontribution, whereas the magnetic part is quite small.\nThe total magnetic entropy of Cu 5-NIPA is 5 Rln 2'\n28:8 J mol\u00001K\u00001, which is only 2 % of the total entropy\nof about 1100 J mol\u00001K\u00001released up to 200 K (the\nlatter is obtained by integrating the temperature depen-\ndence ofCp=T).5\nAt 200 K, the heat capacity of Cu 5-NIPA is still very\nfar from its maximal Dulong-Petit value of Cp= 3RN'\n2767 J mol\u00001K\u00001, because Cu 5-NIPA features a large\nnumber of high-energy phonon modes related to the O{H,\nC{C, C{N, and C{O vibrations. The entropy associated\nwith these modes can be released at very high temper-\natures, only. The complexity of the Cu 5-NIPA struc-\nture and the presence of multiple phonon modes of di\u000ber-\nent nature hinder a quantitative analysis of the speci\fc\nheat data. A non-magnetic reference compound would\nbe ideal to extract the magnetic contribution and ana-\nlyze it in more detail. Unfortunately, such a reference\ncompound is presently not available.\nAt low temperatures, the heat capacity of Cu 5-NIPA\ndoes not fall smoothly to zero. The temperature depen-\ndence ofCp=Tshows an increase below 3 K (see the up-\nper left inset of Fig. 2). The origin of this behavior is\npresently unclear. The increase in Cp=Tcould signify a\nSchottky anomaly or a proximity to the magnetic order-\ning transition. However, our susceptibility data, as well\nas NMR (Sec. III C), rule out any magnetic transition\ndown to 2 K. The heat capacity data are consistent with\nthese observations.\nB. ESR\nThe ESR spectrum of Cu 5-NIPA could be detected\nfrom the lowest measured temperature (5 K) up to 135 K.\nAbove 135 K, the ESR line broadens and eventually be-\ncomes invisible. Between 70 K and 135 K, the spec-\ntra contain only one line that can be \ftted with a sin-\ngle Lorentzian function (1L-Fit) providing the ESR pa-\nrameters, the linewidth \u0001 Bandg-factor,g=h\u0017\n\u0016BBres,\nwhereBresis the resonance \feld. Below 70 K, the sin-\ngle ESR line splits, so that at 40 \u000060 K the \ft is only\npossible with two Lorentzian lines (2L-Fit), whereas at\neven lower temperatures a plethora of narrow lines ap-\npear (Fig. 5). At 5 K, more than 20 lines are visible\nin the spectrum. This complex structure is typical in\nlow-temperature ESR spectra of molecular magnets and\nis related to the hyper\fne coupling.50,51The complexity\nof the signal and the presence of multiple Cu sites and\nanisotropy parameters prevent us from a detailed anal-\nysis of the low-temperature powder spectra. One would\nneed much higher frequencies and, preferably, single crys-\ntals of Cu 5-NIPA, in order to discriminate the multiple\nresonances that are visible below 40 K. So we restrict\nourselves to the analysis of the data above 40 K.\nFig. 6 shows the temperature evolution of the ESR g-\nfactor, linewidth, and line intensities. Above 70 K, the\nobserved powder-averaged values of g'2:2 are typical\nfor Cu2+in the planar oxygen environment.28,50,51Our\nmicroscopic insight into the energy spectrum of the Cu 5\nmolecule (Sec. IV B) suggests that the formation of two\nlines below 70 K is related to two lowest S=1\n2energy\nlevels, which are separated by \u0001 '68 K. Indeed, lines\n1 and 2 show di\u000berent temperature evolution. While the\ndP dB/ (arb.□units)\n1.0 1.1 1.2 1.365□K5□K\nx20\nFiel d□(T)FIG. 5. (Color online) Typical ESR signal of Cu 5-NIPA at\n5 K (solid line) and 65 K (dashed line). For better visibility,\nthe intensity of the high-temperature signal was increased by\na factor of 20.\n2.2\n2.1\n300\n200\n100\n0\n0\n40 60 80 100\nTemperature□(K)120 14012\nIntensity□(arb.□units)Linewidth□(mT)2.31L-fit 2L-fit,□line□1 2L-fit,□line□2\nESR -factorg\nFIG. 6. (Color online) Temperature dependence of the ESR\nsignals:g-factor (top), linewidth (middle), intensity (bot-\ntom). The data at higher temperatures (diamonds) were ob-\ntained by \ftting the spectra with one Lorentzian line (1L-Fit).\nOpen circles and triangles show the data from \ftting the spec-\ntra with two Lorentzian lines (2L-Fit). Representative error\nbars are shown for several data points, only. All intensities\nare scaled to the intensity of line 2 at 65 K.\nintensity of line 2 is reduced upon cooling, the intensity\nof line 1 is growing. In Sec. IV B below, we provide the\nexplicit dependence of the e\u000bective g-tensors for the GS\nand the \frst excited doublet (in terms of the individual\ng-tensors of the three inequivalent Cu sites) in order to6\n0 50 100 150 2000.5\n0.0\n80120160\nFWHM□(kHz)69.5 69.6 69.7/c1090H=□1.6608□T\n/c1090H=□1.6608□T/c110(MHz)69.8 69.94.2□K\n9.9□K\n20□K\n69.8□K\n140□K\n189.8□K\n230.5□K\n70.01.0\nIntensity, /I Imax\n250\nTemperature□(K)\nFIG. 7. (Color online) Top panel: Fourier-transform1H\nNMR spectra measured at di\u000berent temperatures. Arrows\nshow the shifts of the left and right shoulders upon cooling.\nNote that while the right shoulder is moving toward higher\nfrequencies, the left shoulder shows a non-monotonic behav-\nior. Bottom panel: Full width at half-maximum (FWHM)\nof1H NMR spectra plotted as a function of temperature ( T)\nmeasured in an external \feld \u00160H= 1:6608 T.\ndemonstrate the origin of their di\u000berence.\nC.1H NMR\n1. Linewidth\nConsidering the nuclear spin I=1\n2of the1H nu-\ncleus, one expects a single spectral line for each of the\n17 nonequivalent proton sites (see Table III of App. B).\nHowever, these lines merge into a single broad line with\na nearly Gaussian line shape over the whole measured\ntemperature range (Fig. 7, top). The linewidth at 215 K\nis 95 kHz and almost comparable with the linewidth re-\nported for other molecular magnets (see, e.g., Ref. 52).\nThe line position shifts weakly with temperature. The\nbottom panel of Fig. 7 shows the temperature evolu-\ntion of the full width at half-maximum (FWHM) of the\n1H NMR line measured in an applied \feld of \u00160H=\n1:6608 T. It is almost temperature-independent at high\ntemperatures and increases progressively upon cooling\nbelow about 30 K.\nThe shape and width of the1H NMR spectra are gov-\nerned by two main interactions: the nuclear-nuclear dipo-\nlar interactions and the hyper\fne couplings between the\nproton and unpaired electrons at the Cu sites. Therefore,we can write FWHM as:53{55\nFWHM/p\nh\u0001\u00172id+h\u0001\u00172im; (3)\nwhere the broadening due to the nuclear-nuclear dipolar\ninteractions (h\u0001\u00172id) is temperature-independent, while\nthe broadening due to the hyper\fne couplings ( h\u0001\u00172im)\nscales with the local susceptibilities. For each given pro-\nton sitep,\nq\nh\u0001\u00172i(p)\nm\nH'X\njA(p)\nj\u001fj; (4)\nwhereA(p)\njis the dipolar coupling constant between the\nproton and the Cu2+ions at site j, and\u001fjis the lo-\ncal susceptibility. We should note here that, in general,\nthe temperature dependence of \u001fjis di\u000berent from that\nof the bulk susceptibility \u001f(Fig. 2), especially at higher\ntemperatures, and they are also di\u000berent from each other\n(see Fig. 15). The sharp increase in FWHM at low\ntemperatures can be ascribed to the corresponding low-\ntemperature increase in \u001fj.\nThe shallow minimum in FWHM observed around\n70 K is paralleled by the peculiar evolution of the spec-\ntral line (Fig. 7, top). On cooling down, the right side of\nthe line shifts weakly but monotonously to the right, fol-\nlowing the central position of the line. The left side, on\nthe other hand, shows a non-monotonic behavior, shifting\nbackwards for the data at 20 K and below. The origin of\nthis feature (and the minimum in FWHM) can in princi-\nple be traced back to the temperature dependence of the\nlocal Cu2+moments, and indeed such a non-monotonic\nbehavior is shown by the moments on the Cu2 sites. As\nwe discuss in detail below (Sec. IV B and Fig. 15), at\nT\u001c\u0001'68 K, the moments of the two Cu2 sites are\nantiparallel to the \feld, owing to the large negative ex-\nchange \feld exerted by the neighboring Cu1 and Cu3\nsites. These exchange \felds are balanced by entropy at\nsome characteristic temperature T\u0003, at which the Cu2\nmoments turn positive. Our calculations based on the\nactual exchange couplings give T\u0003'38 K (see Fig. 15\nbelow). At even higher temperatures, the Cu2 moments\nattain their paramagnetic Curie-like behavior of isolated\nspins. The contribution to the second moment from the\nCu2 sites is then expected to decrease down to zero and\nthen increase again as we cool down, starting from the\nhigh-temperature paramagnetic to the low-temperature\n\\ferrimagnetic\" S=1\n2state of Cu 5-NIPA.\n2. Wipe-out e\u000bect\nThe wipe-out e\u000bect is common in molecular\nnanomagnets,39,56{58and refers to the gradual loss\nin the NMR signal intensity below a characteristic\ntemperature. Here, we have measured the temperature\ndependence of Mxy(0)T, whereMxy(0) is the transverse7\n10 100/c1090H=□0.9135□T0.6\n0.4\n0.2\n0.00.81.01.2\nIT I T/max max\nTemperature□(K)\nFIG. 8. (Color online) Normalized integrated intensity\n(IT=I maxTmax) measured at \u00160H= 0:9135 T. The strong\ndrop in the intensity below 100 K is the wipe-out e\u000bect. The\nline shows the maximum intensity ( ImaxTmax) attained at high\ntemperatures.\nmagnetization at time t= 0, obtained by the extrapola-\ntion of theMxy(t) recovery curve to t=0. The integrated\nintensityMxy(0)Tis proportional to the total number\nof protons resonating at the irradiated frequency. The\nnormalized integrated intensity of the NMR signal in\nCu5-NIPA as a function of temperature is shown in\nFig. 8. We \fnd that the loss of the NMR signal begins\nbelow 150 K and becomes more pronounced at low\ntemperatures. The onset of the NMR signal loss around\n150 K coincides with the regime where 1 =T1starts\nincreasing towards the maximum (see Fig. 9 below).\nThe loss in the NMR signal is related to the slowing\ndown of the electron spin dynamics which, as we explain\nbelow, is also responsible for the enhancement in 1 =T1.\nOn decreasing T, a progressively larger fraction of pro-\ntons close to the magnetic Cu2+ions attain spin-spin\nrelaxation times ( T2) shorter than the dead time ( \u001cd) of\nthe spectrometer, hence the signal from these protons\ncannot be detected. At low temperatures, only the pro-\ntons that are very far from the magnetic ions contribute\nto the signal intensity. For further details, see Refs. 56\nand 57.\n3. Nuclear spin-lattice relaxation rate 1=T1\nThe spin-lattice relaxation rate 1 =T1for Cu 5-NIPA was\nmeasured at two di\u000berent applied \felds. The recovery of\nthe longitudinal nuclear magnetization after a saturation\npulse was \ftted well by the stretched exponential func-\ntion (top panel of Fig. 9, inset)\n1\u0000M(t)\nM0'A0e\u0000(t=T1)\f; (5)\nwhereM(t) is the nuclear magnetization at a time tafter\nthe saturation pulse, and M0is the equilibrium magne-\ntization. The value of the exponent \ffor both \felds was\n0.4\n0.0\n0\n0 50 100 150 200 250\nTemperature□(K)2460.81.21.6\n1/ (ms )T1/c451\n1/( (emu□ms□K/mol)/c99TT1/c451\n)10 20 30 t(ms)0.010.114.3□K 22□K 43□K\n160□K\n1 ( )//c45M t M0\n100 50 0 150/c1090H=□0.9135□T\n200Orbach□process\n250\nTemperature□(K)1\n0/c119(10 rad/sec)920.9135□T\n1.6608□TFIG. 9. (Color online) Top panel: Spin-lattice relaxation\nrate 1=T1vs.Tmeasured at two di\u000berent values of external\n\feld, 0.9135 T and 1.6608 T. The inset shows the longitudi-\nnal recovery curves at four representative temperatures; solid\nlines are the \fts using Eq. (5). Bottom panel: 1 =(\u001fT1T)\nvs.Tfor two di\u000berent magnetic \felds. The inset shows the\ntemperature-dependent !c(decay rate of the total moment\nSz) extracted from the 1 =T1data; the solid line is the \ft\nusing Eq. (7).\nfound to decrease slowly from 0.95 to 0.7 upon cooling.\nThe deviation of \ffrom unity re\rects the distribution of\nrelaxation rates according to di\u000berent hyper\fne couplings\nbetween the1H nuclei and Cu2+ions.57,59,60Accordingly,\nthe spin-lattice relaxation rate 1 =T1obtained by \ftting\nthe recovery with Eq. (5) provides a value averaged over\nthe whole distribution.\nTheTdependence of 1 =T1is presented in the top panel\nof Fig. 9 for two values of the external \feld. At high tem-\nperatures ( T>\u0018150 K), 1=T1is almostT-independent.\nThis behavior is typical for uncorrelated paramagnetic\nmoments \ructuating fast and at random.61,62At lower\ntemperatures, 1 =T1increases and passes through a max-\nimum atT'40 K. Such a characteristic enhancement\nhas been found in numerous AFM homometallic38and\nheterometallic39rings built of spins s>1\n2, but in spin-1\n2\nsystems it is very rare. The only example known to us\nis the Cu 6magnet with a high-spin S= 3 GS.63,64Sys-\ntems with predominantly AFM couplings and low-spin\nGS (e.g., V 12having anS= 0 GS, Ref. 65) do not show\nsuch a feature in 1 =T1.\nIn homometallic rings, the maximum in 1 =T1essen-\ntially signals the slowing down of phonon-driven spin\n\ructuations.38,40{42,57The equivalence between the spins\n(by virtue of the nearly perfect rotation symmetry of the\nring) allows to express 1 =T1in terms of the spectral den-\nsity of the total magnetic moment Szof the molecule.40\nAs shown numerically40and by a microscopic theory,41,428\nthe phonon-driven decay of Szproceeds independently\nfrom the remaining observables of the problem, which in\nturn leads to a single Lorentzian form for 1 =T1:\n1\nT1'A\u001fT!c(T)\n!2c(T) +!2\nL; (6)\nwhereAis the average square of the transverse hyper\fne\n\feld,!Lis the nuclear Larmor frequency, and !c=1\n\u001c(T)\nis the decay rate of the total moment Sz. The enhance-\nment in the spin-lattice relaxation rate 1 =T1takes place\nwhen!capproaches the order of magnitude of !L.\nThere are two qualitative di\u000berences between Cu 5-\nNIPA and homometallic rings that should be emphasized,\nthough. First, Cu 5-NIPA features a very sparse excita-\ntion spectrum (Sec. IV B), which means that there are\nvery few spin-phonon channels available for relaxation.\nAccording to Ref. 42, this also means that here !cis\nnot expected to show the strong power-law \u0018T3orT4\nbehavior. Instead, one expects a much weaker Tdepen-\ndence, with longer relaxational times in a wide tempera-\nture range.\nThe second di\u000berence is the fact that the Cu 5molecule\ncomprises three inequivalent Cu sites with di\u000berent local\nmagnetizations, and so the above single-Lorentzian for-\nmula for 1=T1is not valid any longer, but instead a multi-\nLorentzian form should be expected on general grounds,\nas in the case of the heterometallic Cr 7Ni ring.66\nGiven the large spin gap, \u0001 '68 K, one could still\nargue in favor of the single-Lorentzian formula of Eq. (6)\nover a wide low- Trange, since the Wigner-Eckart theo-\nrem allows to replace individual spin operators with the\ntotal spin operators times a constant. However, such a\ntreatment would only capture the resonant spin-phonon\ntransitions between the two Zeeman-split levels of the\nGS, but these transitions are prohibited by Kramer's the-\norem. Instead, for temperatures well below the spin gap\n\u0001'68 K, the relaxation dynamics of the spins must\nbe controlled by inter-multiplet Orbach transitions67be-\ntween the GS doublet and the lowest magnetic excitation\nwith the energy \u0001 (see below). Again, this is quite sim-\nilar to Cr 7Ni, with the only di\u000berence that the spin gap\nis much smaller there ( '14 K).66\nWith the above remarks in mind, we may still use the\nabove single-Lorentzian formula for 1 =T1to extract the\ntemperature dependence of !c, but the latter is now a\nrepresentative measure of the relaxation dynamics (as\nprobed by NMR), and not necessarily the relaxation rate\nof the total moment. For simplicity, we have plotted\n1=\u001fT 1Tvs.Tin the bottom panel of Fig. 9 that clearly\nshows a broad maximum re\recting the slowing down of\n\ructuating moments. Following Eq. (6) and the pro-\ncedure outlined in Ref. 38, we extracted !cfrom the\n1=\u001fT 1Tdata at\u00160H= 0:9135 T. The resulting Tde-\npendence of !c(shown in the lower inset of Fig. 9) con-\n\frms that it is much weaker compared to the strong\npower-law dependence found in homometallic rings.38In\nparticular, there is a wide low- Trange over which !c\nfollows the inter-multiplet Orbach relaxation processes\n300\n200\n100\n0\n0\nEnergy□(eV)4CuTotal\n/c454 /c458400500\nDOS□(eV )/c451100200\nTotal\nCu\nO\n0\n0 0.2 /c450.2 /c450.4FIG. 10. (Color online) LDA density of states for Cu 5-NIPA.\nThe shading shows the contribution of Cu orbitals. The inset\nmagni\fes \fve bands at the Fermi level ( E= 0), with the\natomic contributions denoted by the dashed (Cu) and solid\n(O) lines.\nmentioned above. Following Ref. 67,\n!c'3\u00152\u00013\n2\u0019\u0016h4\u001amv5\u00021\ne\u0001=T\u00001; (7)\nwhere\u001am= 2043 kg/m3is the mass density,22vis the\nsound velocity in Cu 5-NIPA (typical value c\u00181500\nm/sec in nanomagnets68), and\u0015stands for the spin-\nphonon coupling energy parameter, which is related e.g.\nto the \ructuating portion of the Dzyalozinskii-Moriya in-\nteractions present in this system. Note that we neglect\nthe Zeeman splitting contribution to the resonance en-\nergy, because it is negligible compared to \u0001. By \ftting\nthe!cdata with Eq. (7), we obtain \u0015=kB'3:9(v=c)5\n2K.\nA more complete quantitative understanding of 1 =T1\n(e.g. the high-temperature behavior and \feld depen-\ndence) must take into account the multi-exponential be-\nhavior of the relaxation, discussed above, but also the\npresence of the wipe-out e\u000bect discussed above, see e.g.\nRef. 66.\nIV. THEORY\nA. Microscopic magnetic model\nTo determine individual exchange parameters in the\nCu5molecule, we calculate the electronic structure of\nthe Cu 5-NIPA compound. This procedure requires reli-\nable crystallographic information, including precise posi-\ntions of all atoms in the unit cell. However, the available\nstructural data21,22are obtained from x-ray di\u000braction\nthat has only limited sensitivity to the positions of hy-\ndrogen atoms. Therefore, we used the literature data as\na starting model and optimized the hydrogen positions,\nwhereas all other atoms were kept \fxed. The equilibrium\npositions of hydrogen are listed in Table III of App. B.\nThe relaxed structure is 18.6 eV/f.u. lower in energy9\n/c450.2\n/c450.40.0\nX M Y Z T R A à Ã0.2\nEnergy□(eV)\nFIG. 11. (Color online) LDA band structure of Cu 5-\nNIPA (thin light lines) and the \ft with the tight-binding\nmodel (thick dark lines). The Fermi level is at zero en-\nergy (white dashed line). The notation of kpoints is as fol-\nlows: \u0000(0;0;0),X(1\n2;0;0),M(1\n2;1\n2;0),Y(0;1\n2;0),Z(0;0;1\n2),\nT(1\n2;0;1\n2),R(1\n2;1\n2;1\n2), andA(0;1\n2;1\n2), where the coordinates\nare given in units of the reciprocal lattice parameters.\nTABLE I. Interatomic distances d(in\u0017A), Cu{O{Cu bridging\nangles'(in deg), hopping parameters ti(in meV), and ex-\nchange integrals Ji(in K) in Cu 5-NIPA. The Jivalues are\nobtained from DFT+ Ucalculations. The AFM contribu-\ntionsJAFM\ni are evaluated as 4 t2\ni=Ue\u000b, whereUe\u000bis the ef-\nfective on-site Coulomb repulsion. The FM contributions are\nJFM\ni=Ji\u0000JAFM\ni.\ndCu{Cu'Cu{O{Cu tiJAFM\niJFM\niJi\nJ13 3.20 107.9 \u00000:054 34\u000089\u000055\nJ12 3.33 114.5 \u00000:128 191\u0000134 57\nJ23 3.51 125.7 \u00000:161 302\u000034 268\nthan the starting model taken from the literature. This\nenergy reduction should be ascribed to the elongation of\nO{H distances that are unrealistically short (about 0.8 \u0017A)\nin the experimental structural data.22\nThe LDA energy spectrum of Cu 5-NIPA (Fig. 10) com-\nprises narrow lines that represent molecular orbitals of\nthe NIPA molecules and the Cu 5(OH) 2unit. The states\nat the Fermi level belong to the narrow Cu 3 dbands with\na sizable admixture of O 2 p. The local environment of Cu\natoms resembles the conventional CuO 4plaquette units\n(four-fold coordination, see Fig. 1, left). Therefore, the\nhighest-lying Cu 3 dbands have predominantly x2\u0000y2\norigin, following the crystal-\feld levels of Cu2+withx\nandyaxes lying in the plane of the CuO 4plaquette.\nThe \fve Cu atoms of the Cu 5molecule (one molecule\nper unit cell) give rise to \fve bands, with the middle\nband crossing the Fermi level (Figs. 10 and 11). This\nspurious metallicity is due to the strong underestimate\nof electronic correlations in LDA. DFT+ Ucalculations\nreveal the robust insulating behavior with a band gap of\nabout 2.2 eV in reasonable agreement with the light-blue\ncolor of the sample.\nHOO\nO O\nOOH\nCu1/c1093-OH\nHCCFIG. 12. (Color online) Cu dx2\u0000y2-based Wannier function\nfor the Cu1 site in Cu 5-NIPA. Note that the ligands (COO\u0000,\n\u00163-OH\u0000, H2O) a\u000bect the positions of oxygen atoms and their\norbitals. However, only the orbitals of four oxygen atoms\ncontribute to the Wannier function.\nTo evaluate the magnetic couplings, we \ft the \fve Cu\nbands with a tight-binding model (Fig. 11) and extract\nthe relevant hopping parameters tiusing Wannier func-\ntions (WFs) based on the Cu dx2\u0000y2orbital character.69\nTheti's are further introduced into an e\u000bective Hub-\nbard model with the on-site Coulomb repulsion Ue\u000b. As\nthe conditions of the half-\flling and strong correlations\n(ti\u001cUe\u000b) are ful\flled, the lowest-lying excitations can\nbe described by a Heisenberg model with the AFM ex-\nchangeJAFM\ni = 4t2\ni=Ue\u000b. The values of tiandJAFM\ni\nobtained from the LDA band structure are listed in Ta-\nble I and provide an overview of the magnetic couplings\nin Cu 5-NIPA. Considering only nearest-neighbor interac-\ntions, we \fnd a strong coupling between Cu2 and Cu3, a\nsomewhat weaker coupling between Cu1 and Cu2, and a\nrelatively weak coupling between Cu1 and Cu3. Long-\nrange couplings are several times smaller than JAFM\n13.\nThe long-range exchanges in the Cu 5molecule are within\n10 K, whereas the interactions between the molecules are\nbelow 3 K (note the experimental estimate of 0.5 K in\nSec. III A). As the long-range couplings are much weaker\nthanJ12,J13, andJ23, we further restrict ourselves to\nthe minimum microscopic model that comprises three\nnearest-neighbor interactions, only.\nPossible FM contributions to the short-range cou-\nplings require that the evaluation of JAFM\ni is supplied\nby an independent estimate of total exchange couplings\nJi. In a so-called supercell approach, total energies of\ncollinear spin con\fgurations are mapped onto the Heisen-\nberg model to yield the Jivalues, as listed in Table I.\nAll nearest-neighbor interactions have sizable FM compo-\nnentsJFM\ni=Ji\u0000JAFM\ni that reduce J12andJ23, whereas\nJ13eventually becomes ferromagnetic. This way, the tri-\nangular units in the Cu 5molecule feature a combination\nof FM interaction J13and AFM interactions J12andJ23\n(Fig. 1, left).\nThe microscopic origin of the magnetic couplings can\nbe understood from the analysis of Cu-based Wannier\nfunctions (Fig. 12). Each WF features the Cu dx2\u0000y210\n1234\n0\n0 100 200 300 400\nTemperature□(K)/c99(10 emu/mol□Cu)/c452\n400 300 200Experiment□(1□T)\nFit: = =62□K,\n/ =3.5J12/c45J\nJ J13\n23 12\n100 08001200\n400\n0/c99/c451(emu/mol□Cu)/c451\nFIG. 13. (Color online) Fit of the magnetic suscepti-\nbility (\u001f) and inverse magnetic susceptibility ( \u001f\u00001, inset)\nwith the exact-diagonalization result for the Cu 5pentamer\n(J12=\u0000J13= 65 K and J23=J12= 3:5).\norbitals together with the 2 porbitals of the surround-\ning oxygen atoms. These oxygen atoms have di\u000berent\nchemical environment and belong to one of the ligands:\n\u00163-OH\u0000, COO\u0000, and H 2O. While the ligands a\u000bect the\npositions of oxygen atoms (note the slight downward dis-\nplacement of the O atoms in COO\u0000groups) and modify\nthe \\shape\" of the porbitals (note the oxygen atom be-\nlonging to the H 2O molecule), they do not provide any\nsizable long-range contributions to the WFs. The short-\nrange nature of the WFs underlies the diminutively small\nlong-range exchange in Cu 5-NIPA. Further microscopic\naspects of the magnetic interactions in Cu 5-NIPA are dis-\ncussed in Sec. V.\nOur microscopic model can be directly compared to the\nexperimental data. Using full diagonalization for the Cu 5\nmolecule, we calculate thermodynamic properties and re-\n\fne theJiparameters as to match the experiment. This\nway, we \ft the experimental magnetic susceptibility curve\nwithJ12=\u0000J13= 62 K and J23= 217 K (so that\nJ12=J23'2\n7), as well as g= 2:22 and the temperature-\nindependent contribution \u001f0=\u00001:7\u000210\u00004emu/mol Cu\n(Fig. 13). The experimental estimates of Jiare in excel-\nlent agreement with the DFT results (Table I), whereas\ntheg-value matches the high-temperature g'2:2 from\nESR (Sec. III B). Regarding the sizable \u001f0, it may origi-\nnate from the diamagnetic gelatin capsule that was used\nin the susceptibility measurement.\nThe combination of the FM coupling J13and AFM\ncouplingsJ12andJ23implies that the spin triangles in\nCu5are non-frustrated. In a classical picture, all three\ncouplings are satis\fed in a con\fguration with Cu3 and\nCu1 spins pointing up and Cu2 spins pointing down, thus\nleading to the total moment of1\n2per molecule. Indeed,\nat low temperatures Cu 5-NIPA shows paramagnetic be-\nhavior with the magnetic moment of1\n2per molecule (\f\f1\n2i\nstate). In high magnetic \felds, the\f\f3\n2iand, eventually,\f\f5\n2istates will be stabilized. The formation of states with\nthe higher magnetic moment can be seen from the sim-\n2\n1\n0\n0\n0 20 40 60\nFiel d□(T)243\nstatic□fiel d□( =□1.8□K) T\nsimulation□( =□1.8□K) Tpulsed□fiel d□( =□1.4□K) T\nMagnetization(arb.□units)\n( /f.u.)/c109B\n1\n23\n2FIG. 14. (Color online) Upper panel: magnetization of\nCu5-NIPA measured in pulsed magnetic \feld (in arb. units).\nBottom panel: magnetization measured in static \feld up to\n14 T and the simulated curve for J12=\u0000J13= 65 K and\nJ23=J12= 3:5 (both in units of \u0016B/f.u.) The arrow denotes\nthe bend in the pulsed-\feld curve that matches the transition\nbetween the\f\f1\n2iand\f\f3\n2istates of the Cu 5molecule.\nulated magnetization curve of Cu 5-NIPA. The\f\f1\n2istate\npersists up to about 50 T, where the magnetization in-\ncreases abruptly until the\f\f3\n2istate is reached.\nWe endeavored to verify this prediction with magneti-\nzation measurements in pulsed \felds. Unfortunately, the\nresults are strongly a\u000bected by dynamic e\u000bects. The typ-\nical duration of the pulse (about 20 \u0016s) is insu\u000ecient to\nchange the magnetic moment of the molecule. Therefore,\nall features are blurred and, moreover, the curve shows\na non-zero slope above 5 T (Fig. 14, top), in contrast to\nthe \rat plateau that is observed in static \feld (Fig. 14,\nbottom). The apparent mismatch between the pulsed-\n\feld and static-\feld data also prevents us from scaling\nthe high-\feld curve. Nevertheless, the bend observed at\n50\u000055 T is likely the signature of the\f\f1\n2i\u0000!\f\f3\n2itran-\nsition and con\frms our expectations.\nB. Exact Diagonalization\nWe are now ready to evaluate the magnetic energy\nspectrum of each Cu 5molecule and deduce its basic prop-\nerties, according to the DFT values of the exchange cou-\nplingsJi. Using a site labeling convention in accord with\nthe site symmetries, the spin Hamiltonian (disregarding\nanisotropies) reads:\nH=J23S3\u0001S220+J12(S1\u0001S2+S10\u0001S20\u0000S3\u0001S110);(8)\nwhere S220\u0011S2+S20,S110\u0011S1+S10, and we have used\nthe relation J13=\u0000J12. In addition to the full SU(2)\nspin rotation group, this Hamiltonian is also invariant\nunder the inversion operation (in real space) through the\ncentral site S3, which maps S3$S3,S1$S10, and11\nS2$S20. So all eigenstates can be characterized by the\ntotal spinS, its projection Malong some quantization\naxisz, and the parity p(even or odd) under inversion.\nGlobal spectral structure. A straightforward numerical\ndiagonalization of Hyields the energy spectrum given in\nTable II. For each level we also provide the good quantum\nnumbersSandp, as well as the expectation values of the\nsquare of the composite spins S220,S3220\u0011S3+S220, and\nS110. The essential \fnding is that the GS is a total spin\nS=1\n2doublet and the lowest excitation (also a S=1\n2\ndoublet) has a very high energy gap \u0001=0 :313J23'68 K.\nThus the Cu 5molecule behaves as a rigid S=1\n2entity\nforT\u001c\u0001, con\frming the previous experimental picture\nfrom magnetization measurements (Sec. III A 1).\nLooking at Table II, we also \fnd that the spectrum is\norganized in three compact groups of states, which are\nseparated by horizontal lines. All states belonging to a\ngiven group show a very similar value of hS2\n3220i, suggest-\ning that the trimer of S3,S2, and S20plays a special\nrole in the physics of Cu 5-NIPA. Moreover, the expecta-\ntion values of S2\n220,S2\n3220, and S2\n110are for all states very\nclose to the formula S(S+ 1) for some integer or half-\nintegerS, suggesting that these composite spins are al-\nmost conserved quantities. Below, we shall demonstrate\nthat all these spectral features as well as the GS proper-\nties can be physically understood quite naturally on the\nbasis of a strong coupling expansion around the limit of\nJ12=jJ13j=0.\nFrom the spectrum, we can also obtain the level-\ncrossing \felds at which the GS of the molecule changes\nfromS=1\n2toS=3\n2and fromS=3\n2toS=5\n2. We \fnd:\ng\u0016BHc1=kB'0:40J23; g\u0016BHc2=kB'1:43J23:(9)\nWithJ23= 217 K and the average g'2:38 (see pre-\nviously), these numbers yield Hc1'54 T (compare to\nFig. 14) and Hc2'194 T.\nGS properties. To probe the nature of the GS, we\ncalculate several experimentally relevant GS expectation\nvalues. We \frst consider how the total S=1\n2moment of\nthe GS (at low T, this moment can be attained by a small\napplied \feld) is distributed among the \fve Cu sites. At\nT\u001c\u0001, we \fnd\nhSz\nii=hSz\nii0\u0002tanh\u0012g\u0016BH\n2kBT\u0013\n; (10)\nwherehSz\nii0are the GS expectation values (in an in-\n\fnitesimal \feld):\nhSz\n3i0=0:1416;hSz\n2i0=hSz\n20i0=\u00000:1415;\nhSz\n1i0=hSz\n10i0=0:3207:(11)\nThe signs of the above numbers are in agreement with the\nclassical ferrimagnetic GS discussed above (Fig. 1, right),\nwhereby the spins S3,S1, and S10are aligned parallel\nto each other, but antiparallel to S2andS20. However\nthe strong deviation (especially for S3,S2andS20) from\ntheir maximum (classical) value of1\n2indicates that thisTABLE II. The exact spectrum of the spin Hamiltonian H\nof Eq. (8) at y=J12=J23=2\n7. For each state we also provide\nits total spin S, its parity punder the real-space inversion\nthrough the central site S3, as well as the expectation values\nofS2\n220,S2\n220, and S2\n220. The horizontal lines di\u000berentiate the\nthree unperturbed manifolds of the strong coupling limit y=0\n(see text and Fig. 16).\ntotal S parity ptotalE=J 23hS2\n220i hS2\n3220i hS2\n110i\n1\n2even\u00001:3289 1.969 0.795 1.969\n1\n2odd\u00001:0157 1.976 0.75 0.024\n3\n2even\u00000:9286 2 0.893 2\n3\n2odd\u00000:1732 0.086 0.879 1.914\n1\n2even\u00000:0395 0.222 1.011 0.222\n1\n2odd 0.3014 0.024 0.75 1.976\n5\n2even 0.5 2 3.75 2\n3\n2odd 0.5303 1.914 3.621 0.086\n3\n2even 0.5714 2 3.607 2\n1\n2even 0.5827 1.809 3.444 1.8087\n0 10 20 30 40 50−0.100.10.20.30.4\nT(K) \n0 50 100 150 200−0.0100.010.02\nT (K) \nCu1\nCu2Cu3Cu3total \nSz\nCu2total \nSz\nT*=38 KCu1\nFIG. 15. (Color online) Temperature dependence of the local\nmoments of the three inequivalent Cu sites in Cu 5-NIPA as\nobtained by exact diagonalization using our ab initio exchange\ncoupling parameters and a \feld of \u00160H= 1:6608 T. The\ninset shows the same results in a di\u000berent scale, so that the\ndeviation from the low- Tbehavior of Eq. (10), shown by the\ndashed lines, is better visible. Also, T\u0003'38 K denotes the\ncharacteristic temperature where the Cu2 moments change\nsign.\nstate has strong QM corrections, as can be directly seen\nfrom the explicit form of the wavefunction (not shown).\nThis feature will be explained on the basis of the strong\ncoupling description developed below.\nThe deviation of the local moments from Eq. (10) at\nhigh temperatures is shown in Fig. 15 at a \feld of 1 :6608\nT. We note in particular the non-monotonic temperature\ndependence of the magnitude of the Cu2 moments, which\nwas discussed in Sec. III C 1 above in relation to the NMR\nlineshape.\nTo further probe the nature of the GS, we consider the\nstrength of various spin-spin correlations eij\u0011hSi\u0001Sji0.12\nWe \fnde31=e310'+0:2122,e32=e320' \u00000:4810,\ne12=e1020' \u0000 0:4299,e120=e102' \u0000 0:2858, and\ne110=e220'+0:2345. These values corroborate the\nabove QM ferrimagnetic picture for the GS, and in addi-\ntion highlight that the bonds with the largest exchange\ncouplingJ23(i.e., the bonds 3-2 and 3-20) exhibit also\nthe strongest correlations. Moreover, using the above\ncorrelations and the values of the exchange couplings,\none \fnds that the J23-bonds contribute about 72% of\nthe total GS energy E0'\u00001:3289J23.\nWe should remark here that the above result e110=\ne220does not arise from the symmetry of the Hamilto-\nnian, but it is an exact property of the particular GS\nonly (see explanation below).\nE\u000bective g-tensors. To make contact with our ESR\n\fndings in Sec. III B we calculate the e\u000bective g-tensors\nof the GS and the \frst excited state in terms of the in-\ndividual g-tensors of the three inequivalent Cu sites. To\nthis end, one needs the equivalent spin operators within\neach multiplet in order to rewrite the Zeeman Hamilto-\nnian as\nX\niB\u0001gi\u0001Si7!B\u0001ge\u000b\u0001S: (12)\nFor the GS, the equivalent operators can be extracted\nfrom Eq. 11 above, which yields\nge\u000b, GS = +0:0708g3\u00000:1415g2+ 0:3207g1:(13)\nSo the deviation from the isotropic g= 2 value is coming\nmainly from that of g1.\nFor the \frst excited state, our exact wavefunctions give\nthe equivalent operators\nSz\n3!\u00001\n3S; Sz\n2;Sz\n20!+0:1647S;\nSz\n1;Sz\n10!+0:002S:(14)\nleading to\nge\u000b, 1st exc. =\u00001\n3g3+ 0:3294g2+ 0:004g1:(15)\nSo here the deviation from the isotropic g= 2 value is\nmainly governed by that of g2andg3. The contribution\nfromg1is extremely small because in the \frst excited\ndoublet the pair of S1andS10forms almost a singlet\n(see Table II and Sec. IV C below).\nThese expressions demonstrate why the e\u000bective g-\ntensors of the two ESR lines appearing below 70 K are\ndi\u000berent.\nC. Strong coupling description\nWe are now going to show that almost all properties\nof the model at y\u0011J12=\u0000J13=2\n7can be adiabatically\ntraced back { even on a quantitative level { to the strong\ncoupling limit y= 0.Global spectral structure. Aty= 0, the spins S1and\nS10are isolated and are, thus, free to point up or down.\nOn the other hand, the spins S3,S2, and S20form an\nAFM spin trimer with the Hamiltonian\nH0=J23=S3\u0001(S2+S20) =1\n2\u0000\nS2\n3220\u0000S2\n220\u0000S2\n3\u0001\n:(16)\nHence the spectrum of H0can be derived analytically in\nterms of the good quantum numbers S220andS3220:\nS220; S3220;parity;deg; E(0)\n11\n2even 8\u0000J23\n01\n2odd 8 0\n13\n2even 16 + J23=2(17)\nwhere we have also indicated the parity and the degen-\neracy (deg) of each unperturbed level. The latter takes\ninto account the four possible states of the space of S1\nandS10, and the Zeeman degeneracy. So we \fnd three\nwell separated unperturbed manifolds. The remaining\ncouplingsV\u0011H\u0000H 0split these manifolds but, as we\nshow below, most of these splittings are fairly weak, thus\nexplaining the overall spectral structure at y=2\n7(Ta-\nble II).\nSince each unperturbed manifold has a well-de\fned\nS3220, we can \fnd the \frst-order splitting with the use\nof the \\equivalent operators\":\nS3!\u00153S3220;S2!\u00152S3220;S20!\u001520S3220;(18)\nwhere the values of \u00151;2;20can be derived from the cou-\npling (Clebsh-Gordan) coe\u000ecients of the respective an-\ngular momenta of each manifold (see below). Using\n\u00152=\u001520, which holds in all manifolds, and replacing (18)\ninV, we obtain\nV!Je\u000bS3220\u0001S110=1\n2Je\u000b\u0000\nS2\u0000S2\n3220\u0000S2\n110\u0001\n;(19)\nwhereJe\u000b= (\u00152\u0000\u00153)y. Thus, to lowest order, the per-\nturbation gives rise to an e\u000bective exchange coupling be-\ntween S110andS3220.\nFor the lowest eight-dimensional manifold, we couple\nthe spinS3=1\n2to the spin S220= 1 to get an S3220=1\n2\nobject, which gives (see Appendix A): \u00153=\u00001\n3,\u00152=\n\u001520= +2\n3, and thusJe\u000b=y>0. The resulting correction\nto the lowest unperturbed manifold reads:\nS220; S3220; S110; S; parityp;deg;\u0001E(1)\n11\n211\n2even 2\u0000y\n11\n201\n2odd 2 0\n11\n213\n2even 4 + y=2(20)\nSimilarly, the second unperturbed manifold has S220=\n0 and thus\u00152=\u001520=0 and\u00153=1, givingJe\u000b=\u0000y, which\nis now FM. So the \frst-order splitting of this manifold has\nthe opposite structure from that of the lowest manifold,13\n/g4/g4/g2/g5 /g4/g2/g6 /g4/g2/g7 /g1/g5/g1/g4/g2/g9 /g4/g4/g2/g9 \n/g20/g10/g12 /g5/g6 /g10/g1/g12/g5/g7 /g11/g16/g14/g18/g15/g20/g3/g12 /g6/g7 \n/g1/g1\n/g4/g4/g2/g5 /g4/g2/g6 /g4/g2/g7 /g1/g4/g2/g6 /g1/g4/g2/g5 /g4/g4/g2/g5 /g4/g2/g6 /g4/g2/g7 /g4/g2/g8 /g4/g2/g9 \n/g20/g10/g12 /g5/g6 /g10/g1/g12/g5/g7 /g4/g4/g2/g5 /g4/g2/g6 /g4/g2/g7 /g1/g4/g2/g9 /g1/g4/g2/g6/g9 /g4/g4/g2/g6/g9 \n/g20/g10/g12 /g5/g6 /g10/g1/g12/g5/g7 /g2/g13/g6/g21/g1/g2/g13/g5/g21/g1/g2/g13/g19/g17/g19/g21/g1\n/g2/g13/g7/g21/g1/g14/g7/g5 \n/g14/g6/g5/g22\n/g14/g5/g6 /g14/g7/g6 /g14/g5/g5/g22(a) Energy spectrum (b) local moments (c) spin-spin correlations \nS=1/2, odd \nS=1/2, even S=3/2, even S=3/2, odd S=5/2, even \nFIG. 16. (Color online) Accuracy of the strong-coupling description. Evolution of the energy spectrum (a), the local moments\n(b), and the spin-spin correlations eij=hSi\u0001Sji(c) from the strong coupling limit y=0 up toy=2\n7. The energies are given in\nunits ofJ23.\nnamely:\nS220; S3220; S110; S; parityp;deg;\u0001E(1)\n01\n213\n2odd 4\u0000y=2\n01\n201\n2even 2 0\n01\n211\n2odd 2 + y(21)\nThe third manifold has S3220=3\n2and thus\u00153=\u00152=\n\u001520=1\n3, which in turn gives Je\u000b= 0. Thus, the third\nunperturbed manifold remains intact in lowest order, i.e.,\nall 16 states have \u0001 E(1)=0:\nS220; S3220; S110; S; parityp;deg;\u0001E(1)\n13\n211\n2even 2 0\n13\n213\n2even 2 0\n13\n215\n2even 4 0\n13\n203\n2odd 4 0(22)\nIn particular, the energy of the state with the maximum\nspin (S=5\n2) remains equal to J23=2 in all orders of\nperturbation theory, as can be easily checked, e.g., for its\nmaximum polarized M=5\n2portion, which is a trivial\neigenstate ofH.\nUp to this lowest order, the level-crossing \felds Hc1\nandHc2are given by\ng\u0016BHc1=kB'3y=2; g\u0016BHc2=kB'(3J23\u0000y)=2:(23)\nWithJ23= 217 K,y= 2J23=7, and the average g'2:2,\nthese numbers give: Hc1'54 T, andHc2'170 T, which\nare close to the exact numbers at y=2\n7given above.\nFigure 16 shows the evolution of the spectrum from\ny= 0 up to y=2\n7, and demonstrates that the above\n\frst-order spectral structure reproduces quantitatively\nthe spectrum at y=2\n7.\nWe should add here that S220,S3220, andS110do not\nremain good quantum numbers in higher orders of theperturbative expansion, which is expected, since the cor-\nresponding angular momenta are not conserved quanti-\nties under the full Hamiltonian. Still, their expectation\nvalues reported in Table II are very close to the above\nstrong-coupling values.\nAs a \fnal remark, we note that one may use the above\nbasis provided by the lowest-order theory and the good\nquantum numbers S,M, andpto \fnd the invariant\nsubspaces ofH. Most of the subspaces turn out to be\ntwo-dimensional, except for the space with S=M=1\n2\nandp= even (in which the GS belongs), which is three-\ndimensional. Here, it may be readily checked that each\nof the three states has S110=S220, thus explaining the\nequalitye110=e220discussed above. The excited states\ndo not share this property, so at higher temperatures the\ntwo correlations deviate from each other.\nGS properties. Let us now return to the nature of the\nGS. According to the above lowest order result, the GS is\na totalS=1\n2doublet with even parity and total energy\nE(1)=\u0000J23\u0000J12. To understand the nature of this state,\nwe note that it arises from the coupling of a spin s=1\n2\n(here S3220) and a spin s= 1 (here S110) to a total S=1\n2\nstate. So we may again use the \\equivalent operators\"\nS3220!\u00001\n3S;S1;S10!+2\n3S; (24)\nwhich in conjunction with the ones relating S1,S3, and\nS5toS3220(see Eq. (18) above) gives the central rela-\ntions:\nS3!+1\n9S;\nS2;S20!\u00002\n9S;S1;S10!+2\n3S:(25)\nThis equation provides a very clear explanation for\nthe peculiar local moment distribution discussed above,\nsince, when the total S=1\n2moment is saturated (by14\napplying a small \feld), this equation gives:\nhSz\n3i= +1\n18;\nhSz\n2i=hSz\n20i=\u00001\n9;hSz\n1i=hSz\n10i= +1\n3:(26)\nThe same numbers can be extracted directly from the\nexplicit form of the GS wavefunction, which for the M=\n+1\n2portion reads:\nj*i=\u00002\n3j\"i3jt1i110jt\u00001i220+1\n3j\"i3jt0i110jt0i220\n+p\n2\n3j#i3\u0000\njt1i110jt0i220\u0000jt0i110jt1i220\u0001\n; (27)\nwherejt1iij=j\"i\"ji,jt\u00001iij=j#i#ji, andjt0iij=\n(j\"i#ji+j#i\"ji)=p\n2. The leading term of Eq. (27)) is\nthe classical ferrimagnetic con\fguration. However, the\noverall form of the wavefunction demonstrates explicitly\nthe presence of strong QM corrections to the GS.\nThe strong coupling limit reproduces the ferrimag-\nnetic arrangement of the local moments and even their\nstrengths, except for S3which has almost tripled its\nmoment at y=2\n7. This is further demonstrated in\nFig. 16(b), which shows the evolution of the local mo-\nments from y=0 up toy=2\n7. All moments, except hSz\n3i,\nexhibit a small slope, i.e., higher-order corrections are\nweak.\nIn a similar fashion, the strong coupling prediction\nfor the spin-spin correlations are: e31=e310=1\n6,e32=\ne320=\u00001\n2,e12=e1020=\u00001\n3,e120=e102=\u00001\n3, and\ne110=e220=1\n4, which are in close agreement with the\nexact values for y=2\n7given earlier. This is again demon-\nstrated in Fig. 16(c) that shows the evolution of the spin-\nspin correlation strengths from y= 0 up toy=2\n7. All\ncorrelations exhibit weak corrections, except e12which\nshows a somewhat larger slope in y.\nV. DISCUSSION AND SUMMARY\nThe Cu 5-NIPA molecule comprises two non-frustrated\nspin triangles with two AFM couplings and one FM\ncoupling each (Fig. 1, left). This coupling regime is\nrather unusual and deserves a further analysis with re-\nspect to the underlying crystal structure and interacting\norbitals. Three leading couplings { J12;J13, andJ23{\nrepresent the Cu{O{Cu superexchange running via the\n\u00163-O atom in the middle of the Cu 3triangle. According\nto Goodenough-Kanamori rules, high values of the bridg-\ning angle should lead to an AFM interaction, whereas low\nbridging angles close to 90\u000efavor FM couplings. Our re-\nsults for Cu 5-NIPA (Table I) follow this general trend.\nHowever, a closer examination pinpoints two peculiari-\nties of this compound.\nFirst, the coupling remains FM for the bridging angle\nof 107:9\u000e, although the standard and commonly accepted\nthreshold value of the FM-AFM crossover is slightly be-\nlow 100\u000e.70Second,JFM\n12is much larger than JFM\n13, eventhough the respective bridging angle is also larger and\nshould lead to a smaller FM contribution. Both pecu-\nliarities should be traced back to the twisted con\fgu-\nration of the interacting CuO 4plaquettes (Fig. 1, left).\nThe systematic work on the angular dependence of the\nexchange coupling is usually restricted to systems with\ntwo plaquettes lying in the same plane or only weakly\ntwisted (the dihedral angle between the planes is close to\n180 deg). The Cu 5molecule represents an opposite limit\nof strongly twisted CuO 4plaquettes, with the dihedral\nangles of 91.8 deg ( J12), 125.2 deg ( J13), and 99.9 deg\n(J23) between the CuO 4planes.71\nThe large FM contribution to J12(JFM\n12=\u0000134 K)\ncan be ascribed to the nearly orthogonal con\fguration\nof the interacting CuO 4plaquettes. However, the mech-\nanism of this FM interaction is yet to be determined.\nThe large FM coupling for the bridging angle of 90 deg\nis generally understood as the Hund's coupling on the\noxygen site72or the direct FM exchange between Cu and\nO.73Both mechanisms depend solely on the Cu{O{Cu\nangle and should be rather insensitive to the mutual ori-\nentation of the CuO 4plaquettes. Therefore, other ef-\nfects, such as the direct Cu{Cu exchange and the Cu{\nO{Cu interaction involving fully \flled Cu 3 dorbitals,\nmay be operative. Experimental information on the mag-\nnetic exchange between the strongly twisted CuO 4pla-\nquettes remains scarce and somewhat unsystematic. The\nCu{O{Cu angle of 104.5 deg combined with the sizable\ntwisting give rise to a FM nearest-neighbor interaction\nin the kagome material kapellasite Cu 3Zn(OH) 6Cl2,74\nwhereas a similar geometry with the Cu{O{Cu angle of\n107.6 deg results in an overall AFM coupling in dioptase,\nCu6Si6O18\u00016H2O.75\nWe also note that Cu 5-NIPA brings a fresh perspec-\ntive on magnetostructural correlations in Cu 3triangular\nmolecules, where the Cu{X{Cu bridging angle was previ-\nously considered the key geometrical parameter. As long\nas the ligand atom X lies in the Cu 3plane, the bridg-\ning angles remain close to 120 deg and should generally\nlead to the AFM exchange. The FM exchange would\nonly be possible when the ligand atoms are shifted out of\nthe plane, as in the Cl- and Br-containing Cu 3molecules\nwhere the Cu{X{Cu angles are below 90 deg.76The twist-\ning of the CuX 4plaquettes provides another opportunity\nfor creating the FM exchange and, moreover, for intro-\nducing it selectively. Surprisingly, organic ligands have\nonly a weak e\u000bect on the magnetic interactions. Al-\nthough carboxyl groups (COO\u0000) provide an additional\nsuperexchange pathway between Cu1 and Cu3 (Fig. 1,\nleft), their molecular orbitals do not in\ruence the Cu\ndx2\u0000y2Wannier functions (Fig. 12). For example, the\nFM contribution to J12exceeds that to J13. This demon-\nstrates that the FM exchange is basically unrelated to the\norganic ligands. It should be understood as a joint e\u000bect\nof the the low Cu{O{Cu angle and twisting.\nThe combination of FM and AFM couplings also has\nan important e\u000bect on the magnetic GS of Cu 5-NIPA.\nA regular spin triangle entails the four-fold degenerate15\nGS that further splits into two close-lying doublets by\nvirtue of anisotropy,27residual interactions between the\ntriangles,10or a marginal distortion of the triangle.28In\nCu5-NIPA, there is only one doublet state, which is sep-\narated by about 68 K from the \frst excited state (see\nTable II). According to Kramer's theorem, in zero \feld\nthe degeneracy of this state can not be lifted, hence the\ntunnel splitting is exactly zero, and no magnetization\nsteps (Landau-Zener-St uckelberg transitions) should oc-\ncur, in contrast to, e.g., the V 6molecule.10Therefore,\nbroad butter\ry hysteresis e\u000bects but without tunneling\nare expected in the magnetization process of Cu 5-NIPA.\nOf particular interest is our experimental \fnding of the\nenhanced1H nuclear spin-lattice relaxation rate 1 =T1at\na characteristic temperature slightly below the spin gap\n(T'40 K). While such an enhancement has been re-\npeatedly found in numerous AFM homometallic38and\nheterometallic39rings of spins S >1\n2, it is very rare for\nS=1\n2and has been observed only in molecules with a\nhigh-spin GS.63,64The origin of the peak is likely the\nsame in both cases, namely, the slowing down of the\nphonon-driven magnetization \ructuations.38,40{42How-\never, the sparse excitation spectrum and the presence\nof nonequivalent Cu sites with di\u000berent local magnetiza-\ntion render Cu 5-NIPA dissimilar to typical homometallic\nrings. We argue that in Cu 5-NIPA inter-multiplet Or-\nbach processes make the dominant contribution to the\nspin-lattice relaxation process, as in the heterometallic\nring Cr 7Ni at very low temperatures.66\nThe use of molecular magnets in quantum computing\nis severely restricted by their short coherence time. Co-\nherence times on the order of 100 \u0016s could be achieved\nin, e.g., V 15by arranging magnetic molecules in a self-\nassembled layer formed by an organic surfactant.18This\nmethod is fundamentally similar to the formation of the\nmetal-organic framework in Cu 5-NIPA. Therefore, it may\nbe interesting to study the decoherence process in this\nmolecular magnet and, more generally, explore the role\nof organic bridges in the decoherence process.\nFinally, we would like to note that only a few examples\nof magnetic pentamers based on spin-1\n2ions have been re-\nported so far. The [Cu( \u0016-L)3]2Cu3(\u00163-OH)(PF 6)3\u00015H2O\n(L\u0000= 3;5-bis(2-pyridil)pyrazolate) shows a somewhat\nsimilar phenomenology with the spin S=1\n2GS and the\nrespective magnetization plateau ranging up to 30 T.77\nHowever, the geometrical structure of this compound is\na trigonal bipyramid that is notably di\u000berent from the\nhourglass shape of the magnetic cluster in Cu 5-NIPA.\nThe magnetic pentamer based on Cu 5(OH) 4(H2O)2(A-\n\u000b-SiW 9O33)2] is more similar to our case and also reveals\nthe spinS=1\n2GS.51,78Nevertheless, the equivalence of\nJ23andJ13, as well as sizable long-range couplings, dis-\ntinguish its spectrum and magnetic properties from that\nof Cu 5-NIPA.\nIn summary, we have studied thermodynamic proper-\nties, spin dynamics, microscopic magnetic model, and en-\nergy spectrum of the spin-1\n2Cu5pentamer in Cu 5-NIPA.\nThis magnetic molecule has an hourglass shape with twoAFM couplings and one FM coupling on each of the two\ntriangles that form the non-frustrated pentamer. In zero\n\feld, the ground state of Cu 5-NIPA is a doublet with the\ntotal spin of S=1\n2. This ground state is separated from\nthe \frst excited state by an energy gap of \u0001 '68 K. Our\nresults evidence a highly inhomogeneous distribution of\nmagnetization over rhe nonequivalent Cu sites accord-\ning to the quantum nature of the magnetic ground state.\nThe maximum in the spin-lattice relaxation rate 1 =T1is\nvery rare among molecular magnets with spin-1\n2ions and\na low-spin ground state.\nACKNOWLEDGMENTS\nRN was funded by MPG-DST (Max Planck\nGesellschaft, Germany and Department of Science\nand Technology, India) fellowship. AT was supported\nby the Alexander von Humboldt Foundation and the\nMobilitas program of the ESF (grant MTT77). OJ\nacknowledges partial support of the Mobilitas grant\nMJD447. IR was funded by the Deutsche Forschungsge-\nmeinschaft (DFG) under the Emmy-Noether program.\nThe high-\feld magnetization measurements were sup-\nported by EuroMagNET II under the EC contract\n228043. Finally, we would like to thank M. Belesi for\nfruitful discussions.\nAppendix A: Coupling of a spin 1 to a spin 1/2\nobject and \\equivalent operators\"\nHere we work out the coupling of two spins, one with\nspinSa= 1=2 and the other one with Sbc= 1. For the\nlatter, we shall imagine that there are two spins Sb=\nSc= 1=2 forming a triplet, namely jt1i=j\"\"i,jt\u00001i=\nj##i, andjt0i=1p\n2(j\"#i+j#\"i). The states of the system\ncan be labeled by jS;Mi, whereSis the total spin (here\nS=1\n2or3\n2), andMis the projection along some axis z.\nThe states with S=3\n2are given by:\nj3\n2;3\n2i=j\"ia\njt1ibc;\nj3\n2;1\n2i=1p\n3\u0010\nj#ia\njt1ibc+p\n2j\"ia\njt0ibc\u0011\n;\nj3\n2;\u00001\n2i=1p\n3\u0010\nj\"ia\njt\u00001ibc+p\n2j#ia\njt0ibc\u0011\n;\nj3\n2;\u00003\n2i=j#ia\njt\u00001ibc:\nOn the other hand, the two states of the S=1\n2doublet\nread:\nj1\n2;1\n2i=1p\n3\u0010\nj\"ia\njt0ibc\u0000p\n2j#ia\njt1ibc\u0011\n;\nj1\n2;\u00001\n2i=1p\n3\u0010\n\u0000j#ia\njt0ibc+p\n2j\"ia\njt\u00001ibc\u0011\n:\nUsing the above explicit relations, it is straightforward to\n\fnd out the equivalent operators within the 2 \u00022 manifold16\nof theS=1\n2doublet. We \fnd:\nSa!\u00001\n3S;Sb!+2\n3S;Sc!+2\n3S: (A1)\nAppendix B: Structural data\nTable III shows the equilibrium positions of the 17 hy-\ndrogen nuclei in the structure.\nTABLE III. Relaxed hydrogen positions in Cu 5-NIPA. Last\ncolumn lists atoms bonded to the given hydrogen atom.\nThe notation of atoms follows Ref. 22, a= 10:8303 \u0017A,\nb= 11:4692 \u0017A, andc= 11:5697 \u0017A.\nx=a y=b z=c\nH2 0.0341 0.9481 0.3828 (C2)\nH4 0.4295 0.6383 0.4380 (C4)\nH6 0.4442 0.0066 0.2381 (C6)\nH10 0.4563 0.4564 0.2062 (C10)\nH12 0.1278 0.8023 0.0923 (C12)\nH13 0.0814 0.3482 0.1695 (O13)\nH14 0.5705 0.7834 0.9955 (C14)\nH14A 0.1105 0.1813 0.9704 (O14)\nH14B 0.0012 0.2993 0.9133 (O14)\nH15A 0.6634 0.2163 0.2633 (O15)\nH15B 0.8008 0.0961 0.2947 (O15)\nH16A 0.9178 0.6522 0.4705 (O16)\nH16B 0.8144 0.5794 0.5639 (O16)\nH17A 0.3996 0.2783 0.2640 (O17)\nH17B 0.4504 0.1667 0.3606 (O17)\nH18A 0.0954 0.1403 0.6165 (O18)\nH18B 0.1939 0.1190 0.6921 (O18)\n\u0003altsirlin@gmail.com\nyi.rousochatzakis@ifw-dresden.de\n1O. Kahn, Molecular Magnetism (UCH, Berlin, 1990);\nD. Gatteschi, R. Sessoli, and J. Villain, Molecular Nano-\nmagnets (Oxford University Press, New York, 2006).\n2L. Gunther and B. Barbara, in Quantum Tunneling of\nMagnetization (Kluwer, Amsterdam, 1995).\n3L. Thomas, F. Lionti, R. Ballou, D. Gatteschi, R. Sessoli,\nand B. Barbara, Nature 383, 145 (1996).\n4A. Chiolero and D. Loss, Phys. Rev. Lett. 80, 169 (1998).\n5W. Wernsdorfer and R. Sessoli, Science 284, 133 (1999).\n6K. L. Taft, C. D. Delfs, G. C. Papaefthymiou, S. Foner,\nD. Gatteschi, and S. J. Lippard, J. Am. Chem. Soc.\n116, 823 (1994); M.-H. Julien, Z. H. Jang, A. Lascialfari,\nF. Borsa, M. Horvati\u0013 c, A. Caneschi, and D. Gatteschi,\nPhys. Rev. Lett. 83, 227 (1999).\n7P. C. E. Stamp, Nature (London) 383, 125 (1996);\nB. Swarzschild, Phys. Today 50, 19 (1997).8L. Bogani and W. Wernsdorfer, Nature Materials 7, 179\n(2008).\n9M. N. Leuenberger and D. Loss, Nature 410, 789 (2001).\n10I. Rousochatzakis, Y. Ajiro, H. Mitamura, P. K ogerler,\nand M. Luban, Phys. Rev. Lett. 94, 147204 (2005).\n11M. Luban, F. Borsa, S. Bud'ko, P. Can\feld, S. Jun, J. K.\nJung, P. K ogerler, D. Mentrup, A. M uller, R. Modler,\nD. Procissi, B. J. Suh, and M. Torikachvili, Phys. Rev.\nB66, 054407 (2002).\n12I. Chiorescu, W. Wernsdorfer, A. M uller, H. B ogge, and\nB. Barbara, Phys. Rev. Lett. 84, 3454 (2000).\n13A. W. E. Dilg, G. Mincione, K. Achterhold, O. Iakovleva,\nM. Mentler, C. Luchinat, I. Bertini, and F. G. Parak, J.\nBiol. Inorg. Chem. 4, 727 (1999).\n14J. K. Shergill, R. Cammack, and J. H. Weiner, J. Chem.\nSoc., Faraday Trans. 87, 3199 (1991).\n15B. Guigliarelli, C. More, A. Fournel, M. Asso, E. C.\nHatchikian, R. Williams, R. Cammack, and P. Bertrand,17\nBiochemistry 34, 4781 (1995).\n16A. Ardavan, O. Rival, J. J. L. Morton, S. J. Blundell, A. M.\nTyryshkin, G. A. Timco, and R. E. P. Winpenny, Phys.\nRev. Lett. 98, 057201 (2007).\n17G. Mitrikas, Y. Sanakis, C. Raptopoulou, G. Kordas,\nand G. Papavassiliou, Phys. Chem. Chem. Phys. 10, 743\n(2008).\n18S. Bertaina, S. Gambarelli, T. Mitra, B. Tsukerblat,\nA. M uller, and B. Barbara, Nature 453, 203 (2008).\n19C. Schlegel, J. van Slageren, M. Manoli, E. K. Brechin,\nand M. Dressel, Phys. Rev. Lett. 101, 147203 (2008).\n20J. Stolze and D. Suter, Quantum computing (Wiley-VCH,\nWeinheim, 2004).\n21Y. Zhao, M. Padmanabhan, Q. Gong, N. Tsumori, Q. Xu,\nand J. Li, Chem. Comm. 47, 6377 (2011).\n22Z.-Y. Liu, J. Chu, B. Ding, X.-J. Zhao, and E.-C. Yang,\nInorg. Chem. Comm. 14, 925 (2011).\n23L. Balents, Nature 464, 199 (2010).\n24M. Trif, F. Troiani, D. Stepanenko, and D. Loss, Phys.\nRev. Lett. 101, 217201 (2008); L. N. Bulaevskii, C. D.\nBatista, M. V. Mostovoy, and D. I. Khomskii, Phys. Rev.\nB78, 024402 (2008).\n25Y. Kamiya and C. D. Batista, Phys. Rev. Lett. 108, 097202\n(2012).\n26B. Georgeot and F. Mila, Phys. Rev. Lett. 104, 200502\n(2010).\n27G. Chaboussant, R. Basler, A. Sieber, S. T. Ochsenbein,\nA. Desmedt, R. E. Lechner, M. T. F. Telling, P. K ogerler,\nA. M uller, and H.-U. G udel, Europhys. Lett. 59, 291\n(2002); Y. Furukawa, Y. Nishisaka, K. I. Kumagai,\nP. K ogerler, and F. Borsa, Phys. Rev. B 75, 220402(R)\n(2007).\n28K.-Y. Choi, Y. H. Matsuda, H. Nojiri, U. Kortz, F. Hus-\nsain, A. C. Stowe, C. Ramsey, and N. S. Dalal, Phys.\nRev. Lett. 96, 107202 (2006); K.-Y. Choi, N. S. Dalal,\nA. P. Reyes, P. L. Kuhns, Y. H. Matsuda, H. Nojiri, S. S.\nMal, and U. Kortz, Phys. Rev. B 77, 024406 (2008).\n29J. Luzon, K. Bernot, I. J. Hewitt, C. E. Anson, A. K. Pow-\nell, and R. Sessoli, Phys. Rev. Lett. 100, 247205 (2008).\n30A. J. Blake, R. O. Gould, C. M. Grant, P. E. Y. Milne,\nS. Parsons, and R. E. P. Winpenny, J. Chem. Soc. Dalton\nTrans. , 485 (1997); E. K. Brechin, A. Graham, P. E. Y.\nMilne, M. Murrie, S. Parsons, and R. E. P. Winpenny,\nPhilos. Trans. Royal Soc. London, Ser. A 357, 3119 (1999).\n31A. M uller, M. Luban, C. Schr oder, R. Modler, P. K ogerler,\nM. Axenovich, J. Schnack, P. C. Can\feld, S. Bud'ko, and\nN. Harison, ChemPhysChem 2, 517 (2001).\n32A. M. Todea, A. Merca, H. B ogge, T. Glaser, J. M.\nPigga, M. L. K. Langston, T. Liu, R. Prozorov, M. Luban,\nC. Schr oder, W. H. Casey, and A. M uller, Angew. Chem.\nInt. Ed. 49, 514 (2010).\n33A. M. Todea, A. Merca, H. B ogge, J. van Slageren,\nM. Dressel, L. Engelhardt, M. Luban, T. Glaser, M. Henry,\nand A. M uller, Angew. Chem. Int. Ed. 46, 6106 (2007).\n34B. Botar, P. K ogerler, and C. L. Hill, Chem. Comm. , 3138\n(2005); A. M uller, A. M. Todea, J. van Slageren, M. Dres-\nsel, H. B ogge, M. Schmidtmann, M. Luban, L. Engelhardt,\nand M. Rusu, Angew. Chem. Int. Ed. 44, 3857 (2005).\n35A. M. Todea, A. Merca, H. B ogge, T. Glaser, L. Engel-\nhardt, R. Prozorov, M. Luban, and A. M uller, Chem.\nCommun. 40, 3351 (2009).\n36I. Rousochatzakis, A. M. L auchli, and F. Mila, Phys. Rev.\nB77, 094420 (2008).37Here,\u00163means that the central oxygen atom is connected\nto three Cu atoms.\n38S. H. Baek, M. Luban, A. Lascialfari, E. Micotti, Y. Fu-\nrukawa, F. Borsa, J. van Slageren, and A. Cornia, Phys.\nRev. B 70, 134434 (2004).\n39H. Amiri, M. Mariani, A. Lascialfari, F. Borsa, G. A.\nTimco, F. Tuna, and R. E. P. Winpenny, Phys. Rev. B\n81, 104408 (2010).\n40P. Santini, S. Carretta, E. Liviotti, G. Amoretti, P. Car-\nretta, M. Filibian, A. Lascialfari, and E. Micotti, Phys.\nRev. Lett. 94, 077203 (2005).\n41I. Rousochatzakis, Phys. Rev. B 76, 214431 (2007).\n42I. Rousochatzakis, A. L auchli, F. Borsa, and M. Luban,\nPhys. Rev. B 79, 064421 (2009).\n43A. A. Tsirlin, B. Schmidt, Y. Skourski, R. Nath, C. Geibel,\nand H. Rosner, Phys. Rev. B 80, 132407 (2009).\n44G. Kresse and J. Furthm uller, Comput. Mater. Sci. 6, 15\n(1996); Phys. Rev. B 54, 11169 (1996).\n45K. Koepernik and H. Eschrig, Phys. Rev. B 59, 1743\n(1999).\n46J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.\nLett. 77, 3865 (1996).\n47O. Janson, I. Rousochatzakis, A. A. Tsirlin, J. Richter,\nY. Skourski, and H. Rosner, Phys. Rev. B 85, 064404\n(2012); A. A. Tsirlin, O. Janson, and H. Rosner, 82,\n144416 (2010).\n48Note that we do not use the temperature-independent term\n\u001f0, because \fts for the pristine sample are done in a narrow\ntemperature range that does not support the evaluation of\nthree variable parameters. The lack of \u001f0results in subtle\ndeviations of the high-temperature e\u000bective moment from\ntheg-value obtained in the model \ft (Fig. 13).\n49For a similar example, see: M. Schmitt, O. Janson,\nM. Schmidt, S. Ho\u000bmann, W. Schnelle, S.-L. Drechsler,\nand H. Rosner, Phys. Rev. B 79, 245119 (2009).\n50B. Cage, F. A. Cotton, N. S. Dalal, E. A. Hillard,\nB. Rakvin, and C. M. Ramsey, J. Amer. Chem. Soc. 125,\n5270 (2003).\n51S. Nellutla, J. van Tol, N. S. Dalal, L.-H. Bi, U. Kortz,\nB. Keita, L. Nadjo, G. A. Khitrov, and A. G. Marshall,\nInorg. Chem. 44, 9795 (2005).\n52C. Maxim, L. Sorace, P. Khuntia, A. M. Madalan, A. Las-\ncialfari, A. Caneschi, Y. Journauxe, and M. Andruha,\nDalton Trans. 39, 4838 (2010).\n53C. P. Slichter, Principle of Magnetic Resonance (Springer-\nVerlag, New York, 1996).\n54A. Abragam, Principles of Nuclear Magnetism (Oxford\nUniversity Press, 1961).\n55M. Belesi, E. Micotti, M. Mariani, F. Borsa, A. Lascialfari,\nS. Carretta, P. Santini, G. Amoretti, E. J. L. McInnes,\nI. S. Tidmarsh, and J. R. Hawkett, Phys. Rev. Lett. 102,\n177201 (2009).\n56M. Belesi, A. Lascialfari, D. Procissi, Z. H. Jang, and\nF. Borsa, Phys. Rev. B 72, 014440 (2005).\n57F. Borsa, A. Lascialfari, and Y. Furukawa (Springer, New\nYork, 2006).\n58P. Khuntia, M. Mariani, A. V. Mahajan, A. Lascialfari,\nF. Borsa, T. D. Pasatoiu, and M. Andruh, Phys. Rev. B\n84, 184439 (2011).\n59P. Khuntia, M. Mariani, M. C. Mozzati, L. Sorace,\nF. Orsini, A. Lascialfari, F. Borsa, C. Maxim, and M. An-\ndruh, Phys. Rev. B 80, 094413 (2009).\n60I. J. Lowe and D. Tse, Phys. Rev. 166, 279 (1968).\n61T. Moriya, Prog. Theor. Phys. 16, 23 (1956).18\n62M. Belesi, X. Zong, F. Borsa, C. J. Milios, and S. P.\nPerlepes, Phys. Rev. B 75, 064414 (2007).\n63A. Lascialfari, D. Gatteschi, A. Cornia, U. Balucani, M. G.\nPini, and A. Rettori, Phys. Rev. B 57, 1115 (1998).\n64S. Carretta, A. Bianchi, E. Liviotti, P. Santini, and\nG. Amoretti, J. Appl. Phys. 99, 08D101 (2006).\n65D. Procissi, A. Shastri, I. Rousochatzakis, M. Al Rifai,\nP. K ogerler, M. Luban, B. J. Suh, and F. Borsa, Phys.\nRev. B 69, 094436 (2004).\n66A. Bianchi, S. Carretta, P. Santini, G. Amoretti, J. Lago,\nM. Corti, A. Lascialfari, P. Arosio, G. Timco, and R. E. P.\nWinpenny, Phys. Rev. B 82, 134403 (2010).\n67A. Abragam and B. Bleaney, Electron paramagnetic res-\nonance of transition ions (Dover publications Inc., New\nYork, 1986) Chap. 10.\n68D. A. Garanin, Phys. Rev. B 78, 020405(R) (2008).\n69H. Eschrig and K. Koepernik, Phys. Rev. B 80, 104503\n(2009).\n70M. Braden, G. Wilkendorf, J. Lorenzana, M. A \u0010n, G. J.\nMcIntyre, M. Behruzi, G. Heger, G. Dhalenne, and\nA. Revcolevschi, Phys. Rev. B 54, 1105 (1996).\n71Here, we averaged four Cu{O vectors of each plaquette,\nbecause Cu and O atoms do not form a single plane owingto the low crystallographic symmetry of the compound.\n72V. V. Mazurenko, S. L. Skornyakov, A. V. Kozhevnikov,\nF. Mila, and V. I. Anisimov, Phys. Rev. B 75, 224408\n(2007).\n73R. O. Kuzian, S. Nishimoto, S.-L. Drechsler, J. M\u0013 alek,\nS. Johnston, J. van den Brink, M. Schmitt, H. Rosner,\nM. Matsuda, K. Oka, H. Yamaguchi, and T. Ito, Phys.\nRev. Lett. 109, 117207 (2012).\n74B. F\u0017 ak, E. Kermarrec, L. Messio, B. Bernu, C. Lhuillier,\nF. Bert, P. Mendels, B. Koteswararao, F. Bouquet, J. Ol-\nlivier, A. D. Hillier, A. Amato, R. H. Colman, and A. S.\nWills, Phys. Rev. Lett. 109, 037208 (2012).\n75O. Janson, A. A. Tsirlin, M. Schmitt, and H. Rosner,\nPhys. Rev. B 82, 014424 (2010).\n76R. Boca, L. Dlh\u0013 a~ n, G. Mezei, T. Ortiz-P\u0013 erez, R. G. Raptis,\nand J. Telser, Inorg. Chem. 42, 5801 (2003).\n77R. Ishikawa, M. Nakano, A. Fuyuhiro, T. Takeuchi,\nS. Kimura, T. Kashiwagi, M. Hagiwara, K. Kindo,\nS. Kaizaki, and S. Kawata, Chem. Eur. J. 16, 11139\n(2010).\n78L.-H. Bi and U. Kortz, Inorg. Chem. 43, 7961 (2004)." }, { "title": "1412.6944v1.Frustration_effects_and_role_of_selective_exchange_coupling_for_magnetic_ordering_in_the_Cairo_pentagonal_lattice.pdf", "content": "Frustration e\u000bects and role of selective exchange coupling for magnetic ordering\nin the Cairo pentagonal lattice\nA. Chainani1,\u0003and K. Sheshadri2,y\n1RIKEN SPring-8 Centre, Sayo-cho, Sayo-gun, Hyogo 679-5148, Japan\n2226, Bagalur, Bangalore North Taluk, Karnataka - 562149, India\n(Dated: June 1, 2022)\nThe Cairo pentagonal lattice, consisting of an irregular pentagonal tiling of magnetic ions on\ntwo inequivalent sites (3- and 4-co-ordinated ones), represents a fascinating example for studying\ngeometric frustration e\u000bects in two-dimensions. In this work, we investigate the spin S= 1=2 Cairo\npentagonal lattice with respect to selective exchange coupling (which e\u000bectively corresponds to a\nvirtual doping of x= 0;1=6;1=3), in a nearest-neighbour antiferromagnetic Ising model. We also\ndevelop a simple method to quantify geometric frustration in terms of a frustration index \u001e(\f;T),\nwhere\f=J=~J, the ratio of the two exchange couplings required by the symmetry of the Cairo lattice.\nAtT= 0, the undoped Cairo pentagonal lattice shows antiferromagnetic ordering for \f\u0014\fcrit= 2,\nbut undergoes a \frst-order transition to a ferrimagnetic phase for \f > \f crit. The results show that\n\u001e(\f;T= 0) tracks the transition in the form of a cusp maximum at \fcrit. While both phases show\nfrustration, the obtained magnetic structures reveal that the frustration originates in di\u000berent bonds\nfor the two phases. The frustration and ferrimagnetic order get quenched by selective exchange\ncoupling, and lead to robust antiferromagnetic ordering for x= 1/6 and 1/3. From mean-\feld\ncalculations, we determine the temperature-dependent sub-lattice magnetizations for x= 0;1=6\nand 1=3. The calculated results are discussed in relation to known experimental results for trivalent\nBi2Fe4O9and mixed valent BiFe 2O4:63. The study identi\fes the role of frustration e\u000bects, the ratio\n\fand selective exchange coupling for stabilizing ferrimagnetic versus anti-ferromagnetic order in\nthe Cairo pentagonal lattice.\nPACS numbers: 81, 75.10.-b, 61.14.-x\nI. INTRODUCTION\nGeometric frustration on a lattice is the inability to\nconsistently satisfy all pair-wise spin interactions be-\ntween lattice sites as de\fned by a Hamiltonian. The most\ncommon example of geometric frustration is the clas-\nsical nearest-neighbour antiferromagnetic (NNAF) Ising\nmodel on a triangular lattice in two dimensions.1This\nseemingly simple case does not show magnetic order\ndown to the lowest temperature, but instead exhibits a\n\fnite entropy at T= 0. Interestingly, the more general\nor complex case, with the Ising-spins replaced by spin\n1/2 Heisenberg-spins has been shown to exhibit an or-\ndered ground state, which corresponds to the quantum\nanalogue of the classical Neel ground state.2This may be\ncompared with another case of frustration-induced zero-\npoint entropy and absence of ordering, namely, the quan-\ntum spin-liquid known for the S= 1=2 kagome lattice\nHeisenberg antiferromagnet.3These contradictory results\ni.e. presence or absence of ordering, typify the uncer-\ntain role of frustration in two dimensional systems with\ntriangular plaquettes. Extensive studies on a variety of\nlattice spin models in two and three dimensions have in-\ndeed revealed and established a plethora of exotic phases\nand properties due to frustration.4{6. Well-known exam-\nples include resonating valence bond superconductivity7,\n'order by disorder'8, spin-ice9, spin liquid with charge\nfractionalization10, magnetic monopoles11, and so on.\nAlthough the triangular plaquette is the smallest unit\nwith intrinsic frustration, any larger plaquette with anodd number of edges (or vertices) would also naturally\ngive rise to frustration. In particular, a pentagonal pla-\nquette also leads to frustration. While a regular pen-\ntagonal plaquette cannot form a Bravais lattice, there\nare 14 known pentagonal tesselations based on irregular\npentagons. Early studies12{14discussed frustration ef-\nfects for the two-dimensional pentagonal lattice obtained\nfrom the hexagonal lattice by cutting each hexagon into\nhalves with parallel lines(Fig. 1 of ref. 12). In the follow-\ning, we call this pentagonal lattice the P1 lattice. Using\na transfer matrix approach, Waldor et al. showed that\nthe NNAF Ising model for the P1 lattice had a \fnite\nground state entropy due to frustration.12For the NNAF\nHeisenberg model on the P1 lattice, Bhaumik and Bose\nidenti\fed the possibility of a collinear Neel type ground\nstate order.13Moessner and Sondhi investigated the P1\nlattice in terms of a NNAF Ising model with a trans-\nverse \feld, and they identi\fed a novel sawtooth state\nconsisting of zigzag antiferromagnetic stripes coexisting\nalternately with frustrated ferromagnetic stripes.14Uru-\nmov, on the other hand, investigated15a di\u000berent tesel-\nlation, the so-called two-dimensional Cairo pentagonal\nlattice (See Fig. 1). Urumov presented an exact solu-\ntion of the nearest-neighbour ferromagnetic Ising model\nfor the two dimensional Cairo pentagonal lattice by map-\nping it onto a Union Jack lattice with nearest and second\nnearest-neighbour non-crossing interactions.15\nHowever, more recently, since the discovery of an ap-\nproximate experimental realization of the Cairo pen-\ntagonal lattice in Bi 2Fe4O9, which exhibits magneto-arXiv:1412.6944v1 [cond-mat.mtrl-sci] 22 Dec 20142\nelectric coupling16and frustration induced non-collinear\nmagnetism,17there has been a signi\fcant resurgence of\ninterest in the Cairo pentagonal lattice.18{24Ralko dis-\ncussed the phase diagram of the XXZ spinS= 1=2\nsystem under an applied magnetic \feld and showed that\nfrustration leads to unconventional phases such as a ferri-\nmagnetic super\ruid.18Rojas et al. employed a direct dec-\noration transformation approach in order to investigate\nantiferromagnetic as well as ferromagnetic coupling for\nan exact solution of the Cairo pentaonal lattice.19They\ncould thereby show that the phase diagram includes a so\ncalled disordered/frustrated state, in addition to ferro-\nmagnetic and ferrimagnetic phases. In an extensive study\nof the NNAF Heisenberg model on the Cairo pentagonal\nlattice, Rousochatzakis et al. revealed the role of an or-\nder by disorder mechanism and a possible spin-nematic\nphase with d-wave symmetry.20The authors addressed\nthe evolution of the phase diagram from the classical to\nthe quantum limit as a function of \f=J=~J, whereJ\nand ~J(J43andJ33, respectively, in ref. 20 ) are the two\ntypes of exchange couplings which originate in the two\ntypes of lattice sites. Jis the exchange coupling between\na 4-co-ordinated and a 3-co-ordinated site, while ~Jis the\nexchange coupling between two 3-co-ordinated sites, re-\nspectively, of the Cairo pentagonal lattice(see Fig. 1(a)\nand its caption).\nOn the experimental front, Retuerto et al. addressed21\nthe role of oxygen non-soichiometry in BiFe 2O5\u0000\u000eand\ncon\frmed that the mixed valence of iron, with a nominal\nvalency of Fe3:2+in BiFe 2O4:63, leads to an antiferro-\nmagnetic ground state with TN= 250K. In the related\nsystem Bi 4Fe5O13F, Abakunov et al. showed22that the\npresence of frustrated exchange couplings lead to a se-\nquence of magnetic transitions at T1= 62 K,T2= 71 K\nandTN= 178 K. Using neutron di\u000braction and thermo-\ndynamic measurements, the authors could show the for-\nmation of a non-collinear antiferromagnet below T1= 62\nK, while the structure between T1andTNwas partially\ndisordered. Pchelkina and Streltsov carried out ab-initio\nband structure calculations23for Bi 2Fe4O9and showed\nthat a complete description required going beyond the\ntwo-dimensional Cairo pentagonal lattice. However, con-\nsidering only the two largest in-plane exchange coupling\nparameters, the obtained ground state was found to be\nconsistent with experiment. And very recently, Nakano\net al. addressed24the magnetization process in the two\ndimensional spin S= 1=2 Heisenberg model on the Cairo\npentagonal lattice. Using a numerical diagonalization\nmethod, they discussed the role of the ratio of the two\ntypes of exchange couplings in driving a quantum phase\ntransition, and very interestingly, found a 1/3 magneti-\nzation plateau usually associated with triangular plaque-\nttes.\nThe results described above clearly show that frustra-\ntion e\u000bects play an important role in determining the\nmagnetic ordering in the Cairo pentagonal lattice. How-\never, the quanti\fcation of frustration and the role of spe-\nci\fc or selective exchange couplings associated with the\nFIG. 1. (Color online) (a) The unit cell of the two-dimensional\nCairo pentagonal lattice(thick blue lines) used in this paper.\nThe undoped ( x= 0) case corresponds to ~J1=~J2=~J, and\nJ1=J0\n1=J2=J0\n2=J, and the magnetic structure depends\non the ratio \f=J/~J. The two magnetic structures obtained\nforx= 0, with \fnite but di\u000berent values of frustration index\n\u001e(T= 0), are shown : (b) for \f= 1, it is antiferromagnetic,\nand (c) for \f= 3, it is ferrimagnetic. For x= 1/6, the ob-\ntained 'star' antiferromagnetic structure is the same as shown\nin (b), while (d) shows the 'boat' antiferromagnetic structure\nobtained for x= 1/3.\ntwo types of sites in the lattice has not been addressed\nyet. Further, while experimental results for the hole-\ndoped system BiFe 2O5\u0000\u000ehave been reported, it is im-\nportant to theoretically investigate the role of doping for\nmagnetic ordering in the Cairo pentagonal lattice. In\nthe present study, we address these issues for the spin\nS= 1=2 Cairo pentagonal lattice in a NNAF Ising model.\nWe \frst develop a method to quantify geometric frustra-\ntion in terms of a frustration index \u001e(\f;T). We \fnd\nthat, for the undoped case (x = 0), \u001e(\f;T= 0) exhibits\na cusp maximum as a function of \f=J=~J. Further,\nin the presence of \fnite frustration, we identify antiferro-\nmagnetic ordering at low \f, which undergoes a \frst-order\ntransition to a ferrimagnetic phase for \f >\fcrit= 2. The\nfrustration and ferrimagnetic order get suppressed by se-\nlective exchange coupling, and lead to antiferromagnetic\nordering for x = 1/6 and 1/3. Mean-\feld calculations\nare carried out to determine the temperature-dependent\nmagnetization for x= 0, 1/6 and 1/3. The results are\ndiscussed in relation to known experimental results for\ntrivalent Bi 2Fe4O9and mixed valent BiFe 2O4:63. The\nstudy identi\fes the role of frustration e\u000bects, the ratio\n\fand selective exchange coupling in relation to ferri-\nmagnetic versus anti-ferromagnetic ordering in the Cairo\npentagonal lattice.3\nII. CALCULATIONAL DETAILS\nThe unit cell is as shown in Fig. 1 (thick blue lines)\nand consists of six Ising spins: four 3-co-ordinated spins\ns1;s2;s3;s4, a 4-co-ordinated spin \u001b0at the center, and\nanother 4-co-ordinated spin \u001b1at the bottom-left corner.\nEach of these can take values of S=\u00061=2. For the most\ngeneral case, the couplings (that are symmetric in the\nindices) are denoted as follows:\nJs1;s3=~J1; Js2;s4=~J2;\nJs1;\u001b0=Js3;\u001b0=J0\n1;\nJs2;\u001b0=Js4;\u001b0=J0\n2;\nJs1;\u001b=Js3;\u001b=J1;\nJs2;\u001b=Js4;\u001b=J2: (1)\nWhile the above detailed notations for the couplings\nare necessary for describing selective exchange couplings,\nwhich e\u000bectively represent virtual doping content ( x), the\nundoped case is obtained by setting ~J1=~J2=~J, and\nJ1=J0\n1=J2=J0\n2=J. As detailed in the following,\nappropriate limits of these parameters allow us to calcu-\nlate physical quantities for x= 1/6 and 1/3. The Ising\nHamiltonian is\nH=X\niHi; (2)\nwhere\nHi=X\n\u000bK(i)\n\u000bs(i)\n\u000b (3)\ncorresponds to the ithunit cell of the two-dimensional\nCairo lattice. Here the symbols K(i)\n\u000bare de\fned by\nK(i)\n1=J1\u001b(i)+J0\n1\u001b0(i)+~J1s(i\u0000y)\n3;\nK(i)\n2=J2\u001b(i+x)+J0\n2\u001b0(i)+~J2s(i+x)\n4;\nK(i)\n3=J1\u001b(i+x+y)+J0\n1\u001b0(i)+~J1s(i+y)\n1;\nK(i)\n4=J2\u001b(i+y)+J0\n2\u001b0(i)+~J2s(i\u0000x)\n2; (4)\nin which the superscripts i\u0006x; i\u0006y; i +x+yare\nindices of unit cells neighboring i. We perform a mean-\n\feld decoupling of the Hamiltonian Eq.(2) according to\nSiSj'hSiiSj+hSjiSi\u0000hSiihSji (5)\nfor a product of any two spins SiandSj, wherehSi=\nTr(Se\u0000HMF=kBT)=Tr(e\u0000HMF=kBT) for anyS. This ap-\nproximation linearizes the unit cell Hamiltonian Eq.(3)\nin the spin variables:\nHMF=c0+4X\n\u000b=1c\u000bs\u000b+c5\u001b0+c6\u001b: (6)\nFIG. 2. (Color online) Plots of (a) frustration index \u001e(\f;0)\n(Eq. (14)), (b) zero-temperature free energy f(\f;0) (Eq.\n(10)), and (c) Emin(\f) (Eq. (13)) as a function of \f=J=~Jfor\nthe undoped case, x= 0 ;\u001e(\f;0) exhibits a non-monotonic\nbehaviour and a cusp at \f= 2.\nWe have suppressed the unit cell index i. The coe\u000ecients\nare\nc0=\u00002~J1m1m3cos(ky)\u0000J1m3m6cos(kx+ky)\n\u00002~J2m2m4cos(kx)\u0000J1m1m6\n\u0000J2m2m6cos(kx)\u0000J2m4m6cos(ky)\n\u0000(J0\n1m1+J0\n1m3+J0\n2m2+J0\n2m4)m5;\nc1= 2~J1m3cos(ky) +J1m6+J0\n1m5;\nc2= (2 ~J2m4+J2m6) cos(kx) +J0\n2m5;\nc3= 2~J1m1cos(ky) +J1m6cos(kx+ky) +J0\n1m5;\nc4= 2~J2m2cos(kx) +J2m6cos(ky) +J0\n2m5;\nc5=J0\n1m1+J0\n1m3+J0\n2m2+J0\n2m4;\nc6=J1m1+J2m2cos(kx) +J1m3cos(kx+ky)\n+J2m4cos(ky): (7)\nFor the sub-lattice magnetizations, we have introduced\nthe notation m\u000b=hs\u000bi(for\u000b= 1;2;3;4),m5=\nh\u001b0i; m 6=h\u001bi. Since the single unit-cell Hamiltonian\n(Eq.(3)) involves couplings with spins in the neighboring\nunit cellsi\u0006x\u0006y, we have introduced a spin density wave\n(SDW) vector kwhose components kx= 2\u0019=\u0015 1; ky=\n2\u0019=\u0015 2appear in the expressions for the coe\u000ecients. The\nsix magnetizations m\u000b(\u000b= 1;\u0001\u0001\u0001;6), and the two wave\nlengths\u00151;\u00152together form an eight-component mean-\n\feld order parameter vector \u0016:\n\u0016= (m1; m 2; m 3; m 4; m 5; m 6; \u00151; \u00152):(8)\nThe thermodynamics of the model is determined by the\nunit cell free energy f(T):\ne\u0000f(T)=kBT=Tr(e\u0000HMF=kBT); (9)4\nFIG. 3. (Color online) Components of the order parameter\n\u0016at zero temperature plotted as a function of \fatx= 0.\n(a) Plots of the sub-lattice magnetizations m1tom6. The\njumps seen for m2,m4andm5indicate a change in the mag-\nnetic order from an antiferromagnetic phase for \f\u00142 to a\nferrimagnetic phase for \f > 2. (b) The SDW lengths \u00151;\u00152\nplotted as a function of \findicate a doubling of the magnetic\nlattice along x and y axes for the antiferromagnetic phase\nfor\f\u00142(panel c) compared to the ferrimagnetic phase for\n\f >2(panel d).\nwhere the trace is performed over the six spins of the unit\ncell, each of which takes the values \u00061=2. Therefore\nf(T) =c0\u00006kBTln(2)\u0000kBT6X\n\u000b=1ln\u0014\ncosh\u0012c\u000b\n2kBT\u0013\u0015\n:\n(10)\nThe magnetizations are determined using the self-\nconsistency equations\n2m\u000b+ tanh\u0012c\u000b\n2kBT\u0013\n= 0; \u000b = 1;\u0001\u0001\u0001;6; (11)\nFIG. 4. Sub-lattice magnetizations mi;i= 1\u00006, plotted as a\nfunction of temperature for the undoped case ( x= 0), for (a)\n\f= 1 (antiferromagnetic phase) and (b) \f= 3 (ferrimagnetic\nphase).\nwhile the SDW lengths are determined by free energy\nminimization @f=@\u0015k= 0; k= 1;2:\n@c0\n@\u0015k+6X\n\u000b=1m\u000b@c\u000b\n@\u0015k= 0; k = 1;2: (12)\nWe solve the eight equations in (11) and (12) numerically\nto obtain the order parameter vector \u0016(fJijg;T).\nWe de\fne a frustration index \u001e(\f;T) in the follow-\ning manner. Consider a reference state for which each\nbond between neighboring spins were to be fully satis-\n\fed. Then, Eq.(3) would give a zero temperature energy\nof\nEmin=\u00001\n2(J1+J0\n1+J2+J0\n2+~J1+~J2): (13)\nThe zero temperature free energy f(\f;T = 0) deviates\nfromEmin(\f) because of frustration, so if we de\fne\n\u001e(\f;T) = 1\u0000f(\f;T)\nEmin(\f); (14)5\nFIG. 5. Sub-lattice magnetizations mi;i= 1\u00006, plotted as\na function of temperature for the antiferromagnetic phases of\nx= 1=6) for (a)\f= 1 and (b) \f= 3.\nthen\u001e(\f;T = 0) is a convenient measure of frustration.\nIt provides a quantitative measure of frustration as a\nfunction of \fand selective exchange coupling.\nIII. RESULTS AND DISCUSSIONS\nWe \frst compute the order parameter vector \u0016by nu-\nmerically solving the eight equations in (11) and (12)\nfor the undoped case with ~J1=~J2=~J, andJ1=\nJ0\n1=J2=J0\n2=J. In Fig. 2, we plot the frustration\nmeasure\u001e(\f;T = 0) (Eq. (14)), the zero-temperature\nfree energy f(\f;T = 0) (Eq. (10)), and Emin(\f) (Eq.\n(13)) as a function of \f=J=~J. The frustration measure\n\u001e(\f;T = 0) exhibits a clear cusp maximum at \f= 2.\nThe cusp originates in the zero-temperature free energy\nf(\f;T= 0) which exhibits a derivative discontinuity at \f\n= 2. This indicates a \frst-order transition as a function\nof\f. As shown in Fig. 3, the order parameter plots as\na function of \findicate an antiferromagnetic phase for\n\f\u00142, which transforms into a ferrimagnetic phase for\n\f >2.\nFIG. 6. Sub-lattice magnetizations mi;i= 1\u00006, plotted as\na function of temperature for the antiferromagnetic phases of\nx= 1=3) for (a)\f= 1 and (b) \f= 3.\nWe see discontinuities in the sub-lattice magnetizations\nm2,m4andm5at\f= 2: while m2andm4\rip from\n1=2 to\u00001=2,m5responds by changing from 0 to 1 =2 to\nminimize the free energy (Fig. 3(a)). The SDW lengths\n\u00151;\u00152also have discontinuities (Fig. 3(b)): \u00151;\u00152= 2\nfor\f\u00142, and\u00151;\u00152= 1 for\f >2. Thus, the magnetic\nunit cell of the antiferromagnetic phase is doubled along\nthe x and y axes, compared to the ferrimagnetic unit cell\nas shown in Fig. 3(c) and (d), respectively. Interestingly,\nwhile the system remains frustrated in both these phases,\nthe frustration index \u001e(T= 0)!0 as\f!0, but goes to\na \fnite value at large \fasymptotically. Thus, in the limit\n\f= 0, we have isolated dimers on the 3-co-ordinated sites\nwhile for\f=1, we retain the ferrimagnetism.\nThe ferrimagnetic phase obtained for \f >2 with a to-\ntal magnetization M(T= 0) = 1=3, is identical to earlier\nstudies from (i) exact calculations for the Ising15,19, (ii)\nhard-core bosons18and (iii) the Heisenberg20models. It\nis noted that, for the antiferromagnetic phase, the sub-\nlattice magnetization m5=h\u001b0i= 0 for\f\u00142, (see Fig.\n3(a)) due to the fact that out of its 4 nearest-neighbour\nsites, two sites have up-spins and two have down-spins.6\nThis leads to an e\u000bective cancellation of the magnetiza-\ntion contribution from the \u001b0site. In earlier work, Rojas\net al.19reported a frustrated phase of Ising spins from ex-\nact calculations, corresponding to the antiferromagnetic\nphase obtained from our mean-\feld calculations, while\nRousochatzakis et al.20obtained a so called orthogonal\nphase for the Heisenberg case.\nWe now turn to investigating the role of selective ex-\nchange coupling at T= 0. We carry out this exercise\nmainly to identify the role of speci\fc couplings of the 3-\nand 4-co-ordinated sites in the Cairo lattice, and to e\u000bec-\ntively simulate virtual dopings of x = 1/6 and 1/3. From\nthe way we have de\fned our couplings (see Eq. (1)), it\ncan be observed that if we set J0\n1=J0\n2= 0, the central\nspin\u001b0gets disconnected from the lattice. Consequently,\nthe central spin \u001b0does not participate in the magnetic\nordering and the system e\u000bectively represents a virtual\nhole doping content of x= 1=6. In the same way, it can\nbe observed that when J2=~J2=J0\n2= 0, the spins s2\nands4get disconnected, corresponding to a virtual hole\ndoping ofx= 1=3. Our results indicate that for both\nthese cases, the system is antiferromagnetic for all \f, i.e.\n\u00151;\u00152= 2. Forx= 1=6, the sub-lattice magnetizations\nare exactly the same as those for x= 0;\f\u00142 (Fig. 3),\nand we therefore refrain from showing these plots in a\n\fgure. For x= 1=3, the actual magnetic structure gets\nmodi\fed as discussed below. Another important di\u000ber-\nence from the x= 0 case is that \u001e(\f;0) = 0 for both\nvalues ofx= 1=6 and 1=3 and is independent of \f, indi-\ncating the absence of frustration.\nThe absence of a \f-driven transition for the cases\nx= 1=6;1=3 is easy to understand. In the x= 1=6\ncase, the central spin \u001b0does not participate in the mag-\nnetic ordering. The remaining spins can be looked upon\nas forming a \\star\" con\fguration with 12 bonds on its\nperiphery, all of which can be satis\fed in the antifero-\nmagnetic phase, thus removing frustration fully. An in-\ncrease of\fonly strengthens the antiferromagnetic order,\nincreasingTN(see Fig. 7 and associated description). In\nthex= 1=3 case, the spins \u001b2;\u001b4do not participate in the\nmagnetic ordering. The remaining spins can be looked\nupon as forming a \\boat\" con\fguration with 10 bonds\non its periphery, all of which can be satis\fed in the antif-\neromagnetic phase, again fully removing frustration(Fig.\n1(d)). An increase of \fin this case also, merely strength-\nens the antiferromagnetic order by increasing TN. While\nboth thex= 1=6 andx= 1=3 cases exhibit antifer-\nromagnetic order, the actual spin ordering is found to\nbe di\u000berent: the former corresponds to a repetition of\n\\star\"(Fig. 1(b)) unit cells, and the latter, a repetition\nof \\boat\"(Fig. 1(d)) unit cells.\nIn Figs. 4(a) and 4(b), we plot the sub-lattice magne-\ntizationsmi;i= 1;\u0001\u0001\u0001;6 for the undoped case ( x= 0),\nas a function of temperature in units of ~Jfor\f= 1\nand\f= 3, i.e. in the antiferromagnetic and ferrimag-\nnetic phases, respectively. We \fnd that the transition\ntemperature TCandTN(the ferrimagnetic and antifer-\nromagnetic transition temperatures, respectively) for de-\nFIG. 7. (a) Plots of the transition temperature Tc;TNas a\nfunction of \ffor x = 0, 1/6 and 1/3. A jump in the ordering\ntemperature is seen only for x = 0 at \f= 2. In (b) we present\na magni\fed view of the region near Tc.\nstruction of magnetic order primarily depends on \f. This\nis further borne out by the temperature-dependence plots\nforx= 1=6 presented in Figs. 5(a) and 5(b) for \f= 1 and\n\f= 3, respectively. The temperature-dependence plots\nforx= 1=3, presented in Figs. 6(a) and 6(b) for \f= 1\nand\f= 3, respectively, also re\rect this. In particular,\nTNincreases on increasing \fforx= 1=6 and 1=3. In\nfact, the ordering temperatures increase smoothly with\n\fas can be seen in Fig. 7(a), where we plot TCand\nTNas a function of \ffor the three cases x= 0;1=6;1=3.\nThe discontinuity at the \frst-order transition for x= 0\nat\f= 2 is quite clear and remarkable. In Fig. 7(b), we\npresent a magni\fed view of the plot around \f= 2.\nWe can also see in Fig. 7(a) that the dependence of\nTNonxis negligible for all \f\u00142. But for \f > 2,\nwhile thex= 1=6;1=3 plots follow the same course as\n\f\u00142, thex= 0 curve follows a completely di\u000berent\ncourse because of the discontinuity. The larger TNof7\nthe antiferromagnetic phase compared to the TCof the\nferrimagnetic phase, just above \f= 2, suggests a higher\nstability of the antiferromagnetic phase. However, for\n\fvalues greater than \u00182:5, the ferrimagnetic TCfor\nx= 0 becomes larger than the antiferromagnetic TNfor\nx= 1=6 and 1=3. The results also show that TNgoes to\na \fnite value as \f!0, but increases linearly for high \f.\nWe now describe the physical picture of the vari-\nous phases obtained by varying the model parameters.\nFirstly, the crucial role that frustration plays in the \f-\ndriven \frst-order transition is seen from the results of the\nx= 0 case. In the x= 0 antiferromagnet region obtained\nfor\f\u00142, the frustration is con\fned to the core of the\nunit cell, i.e. two out of the four J0-bonds (either the\nJ0\n1or theJ0\n2bonds) connecting to the central spin \u001b0are\nfrustrated; in this situation, the frustration measure \u001e(0)\nis an increasing function of \f(see Fig. 2, \f\u0014\fcrit= 2 ).\nWith further increase in \fbeyond\fcrit= 2, it becomes\nenergetically favorable to \\eject\" the frustration from the\ncore to the ~Jbonds lying on the periphery of the unit\ncell i.e. the the four ~J-bonds connecting 3-co-ordinated\nsites become frustrated. Thus, the \f-driven transition is\nattended by a change in the location of the frustrated\nbonds in the unit cell. Secondly, it is surprising to \fnd\nthat the obtained TNvalues do not depend on xbut only\non\f. However, experimental results have also indicated\nthatTNdoes not depend signi\fcantly on doping content.\nFor example, studies on Bi 2Fe4O9(\u0011BiFe 2O4:5; nominal\nvalency of Fe3:0+), the material recently recognized as\na realization of the Cairo pentagonal lattice, reported a\nTN= 238 K for single crystals17,25, while for polycrys-\ntals,TNwas reported to be \u0018260 K.16,26,27For mixed\nvalent polycrystalline BiFe 2O4:63with a nominal valency\nof Fe3:2+, which corresponds to a hole doping of \u001820%,\nRetuerto et al. reported a value of TN= 250 K.21Fur-\nther, forx= 1=6 and 1=3, since the magnetic structures\nhave no frustration, the zero temperature free energy\nf(T= 0) =Emin. However, the two magnetic structures\nhave di\u000berent values of Emin. From Eq.(3), we obtain\nEmin=\u0000(J+~J) forx= 1=6, andEmin=\u0000(J+~J=2 for\nx = 1/3. Thus, even with di\u000berent values of the ground\nstate energies for x= 1=6 and 1=3, our results suggest\nthatTNdepends only on \f=J=~J.\nFinally, the present results also show that the ab-\nsolute values of the sub-lattice magnetizations at high\ntemperatures(\u00180:25\u00000:5TN/TCtoTN/TC)) depend on\nthe co-ordination of the sites and the value of \f. For ex-\nample, as can be seen in Fig. 4(a) for x= 0 and\f= 1,\nthe absolute values of sub-lattice magnetizations of the\n3-co-ordinated sites mi;i= 1;\u0001\u0001\u0001;4 are exactly the same,\nbut for\u00180:5TNtoTN, they are slightly lower than m6,\nwhich is a 4-co-ordinated site. Similarly, for the ferrimag-\nnetic phase with \f= 3, Fig. 4(b) shows that absolutevalues ofmi;i= 1;\u0001\u0001\u0001;4(3-co-ordinated sites) are exactly\nthe same, but for \u00180:25TCtoTC, they are lower than\nm5andm6which are 4-co-ordinated sites. This was also\npointed out by Urumov for the ferrimagnetic phase us-\ning an exact calculation, although the changes were very\nsmall.15Forx= 1=6, the changes are the same as for\nx= 0 and\f= 1, but the di\u000berence between 3 and 4-co-\nordinated sites get enhanced on increasing \f= 3. Very\ninterestingly, we see the opposite behaviour for x= 1=3\ncompared to x= 1=6. For\f= 1(Fig. 6(a)), due to selec-\ntive exchange coupling, the structurally 4-co-ordinated\nsites become magnetically 2-co-ordinated sites(see Fig.\n1(d)), and consequently, the absolute values of m5and\nm6get suppressed compared to m1andm3(structurally\nand magnetically 3-cordinated sites) for temperatures be-\ntween\u00180:25TNtoTN. In contrast, for \f= 3, the dif-\nference between absolute values of m5,m6compared to\nm1,m3become smaller as the magnetization pro\fles get\ndominated by the larger value of Jcompared to ~J.\nIV. CONCLUSIONS\nIn conclusion, we have investigated the spin S= 1=2\nCairo pentagonal lattice with respect to selective ex-\nchange coupling, in a nearest-neighbour antiferromag-\nnetic Ising model. We have developed a simple method\nto quantify geometric frustration in terms of a frustration\nindex\u001e(\f;T), where\f=J=~J, the ratio of the two ex-\nchange couplings required by the symmetry of the Cairo\nlattice. At T= 0, the undoped Cairo pentagonal lattice\nshows a \frst order phase transition with antiferromag-\nnetic order for \f\u0014\fcrit= 2, which transforms to a\nferrimagnet for \f > \fcrit.\u001e(\f;T = 0) exhibits a cusp\nmaximum at \fcrit. The obtained magnetic structures\nreveal that the frustration originates in di\u000berent bonds\nfor the two phases. The frustration and ferrimagnetic\norder get suppressed by selective exchange coupling, and\nthe system shows antiferromagnetic ordering for a virtual\ndoping ofx= 1/6 and 1/3. From mean-\feld calculations,\nwe obtained the temperature-dependent sub-lattice mag-\nnetizations for x= 0;1=6 and 1=3. The calculated results\nwere discussed in relation to known experimental results\nfor trivalent Bi 2Fe4O9and mixed valent BiFe 2O4:63. The\nresults show the fundamental role of frustration and se-\nlective exchange coupling in determining the kind of spin\nordering and how they transform in the Cairo pentagonal\nlattice.\nV. ACKNOWLEDGEMENT\nWe sincerely thank Professor Viktor Urumov, Institute\nof Physics, Macedonia, for sending us Reference 15.\n\u0003chainani@spring8.or.jpykshesh@gmail.com8\n1G. H. Wannier, Phys. Rev. 79, 357(1950).\n2B. Bernu, C. Lhuillier, and L. Pierre, Phys. Rev. Lett. 69,\n2590(1992).\n3S. Depenbrock, I.P. McCulloch, and U. Schollwock, Phys.\nRev. Lett. 109, 067201 (2012)\n4L. Balents, Nature 464, 199(2010).\n5Introduction to Frustrated Magnetism : Materials, Exper-\niments, Theory edited by C. Lacroix, P. Mendels and F.\nMila(Springer, Heidelberg, 2011).\n6C. Castelnovo, R. Moessner, S. L. Sondhi, Annual Review\nof Condensed Matter Physics 3, 35-55 (2012).\n7P. W. Anderson, Science 235, 1196(1987) ; G. Baskaran,\nZ. Zou, and P.W. Anderson Solid State Communications\n63, 973(1987).\n8J. Villain, R. Bidaux, J. P. Carton, and R. J. Conte, J.\nPhys.(Paris), 41, 1263 (1980).\n9A. P. Ramirez, A. Hayashi, R. J. Cava, R. Siddharthan,\nand B. S. Shastry, Nature (London) 399, 333 (1999).\n10P. Fulde, K. Penc, and N. Shannon, Ann. Phys. 11,\n892(2002) ; F. Pollman, P. Fulde and E. Runge, Phys.\nRev. B 73, 125121 (2006).\n11C. Castelnovo, R. Moessner, S. L. Sondhi, Nature 451, 42-\n45 (2008).\n12M. H. Waldor, W. F. Wol\u000b, and J. Zittartz, Z. Phys. B 59,\n43 (1985).\n13U. Bhaumik and I. Bose, Phys. Rev. B 58, 73(1998).\n14R. Moessner and S. L. Sondhi, Phys. Rev. B 63, 224401\n(2001).\n15V. Urumov, J. Phys. A:Math. Gen. 35, 7317 (2002).16A. K. Singh, S. D. Kaushik, B. Kumar, P. K. Mishra, A.\nVenimadhav, V. Siruguri and S. Patnaik, Appl. Phys. Lett.\n92, 132910(2008).\n17E. Ressouche, V. Simonet, B. Canals, M. Gospodinov, and\nV. Skumryev Phys. Rev. Lett. 103, 2672204(2009).\n18A. Ralko, Phys. Rev. B 84, 184434 (2011).\n19M. Rojas, O. Rojas, S. M. de Souza, Phys. Rev. E 86,\n051116 (2012).\n20I. Rousochatzakis, A. M. Lauchli, R. Moessner Phys. Rev.\nB 85, 104415 (2012).\n21M. Retuerto, M. J. Martinez-Lope, K. Krezhov, T.\nRuskov, I. Spirov, P. Krystev, E. Jimenez-Villacorta, M.\nT. Fernandez-Diaz, and J. A. Alonso, Phys. rev. b 85,\n174406(2012).\n22A. M. Abakunov, D. Batuk, A. A. Tsirlin, C. Prescher,\nL. Dubrovinsky, D. V. Sheptyakov, W. Schnelle, J. Hader-\nmann, G. Van Tendeloo, Phys. Rev. B 87, 024423(2013).\n23Z. V. Pchelkina and S. V. Streltsov, Phys. Rev. B 88,\n054424 (2013).\n24H. Nakano, M. Isoda, and T. Sakai, J. Phys. Soc. of Japan\n83, 053702(2014).\n25D. M. Gianchinta, G. C. Papefthymiou, W. M. Davis, and\nH. C. Zur Loye, H. Solid State Chem. 99, 120(1992).\n26N. Shamir and E. Gurewitz, Acta Crystallogr. SEct. A 34,\n662 (1978).\n27A. C. Tutov, I. E. Mylnikova, N. N. Parfenova, V. A.\nBokov, and S. A. Kizhaev, Sov. Phys. Solid State 6,\n963(1964)." }, { "title": "2203.15460v1.Realistic_micromagnetic_description_of_all_optical_ultrafast_switching_processes_in_ferrimagnetic_alloys.pdf", "content": "1 \n Realistic micromagnetic description of all -optical ultrafast switching \nprocesses in ferrimagnetic alloys \n \nV. Raposo1,*, F. García- Sánchez1, U. Atxitia2, and E. Martínez1,+ \n \n1. Applied Physics Department, University of Salamanca. \n2. Dahlem Center for Complex Quantum Systems and Fachbereich Physik . \n*,+: Corresponding author s: victor@usal.es , edumartinez@usal.es \n \nAbstract \nBoth helicity -independent and helicity -dependent all -optical switching processes driven \nby single ultrashort laser pulse ha ve been experimentally demonstrated in ferrimagnetic \nalloys as GdFeCo. Although the switching has been previously reproduced by atomistic \nsimulations, the lack of a robust micromagnetic framework for ferrimagnets limits the \npredictions to small nano -systems, whereas the experiments are usually performed with \nlasers and samples of tens of micrometers. Here we develop a micromagnetic model based \non the extended Landau- Lifshitz -Bloch equation, which is firstly validated by directly \nreproducing atomistic results for small samples and uniform laser heating. After that, the \nmodel is used to study ultrafast single shot all -optical switching in ferrimagnetic alloys \nunder realistic conditions. We find that the helicity -independent switching under a \nlinearly polarized laser pulse is a pure thermal phenomenon, in which the size of inverted \narea directly correlates with the maximum electron temperature in the sample . On the \nother hand, the analysis of the helicity -dependent processes under circular polarized \npulses in ferrimagnetic alloys with different composition indicates qualitative differences \nbetween the results predicted by the magnetic circular dichroism and the ones from \ninverse Faraday effect . Based on these predictions, we propose experiments that would \nallow to resolve the controversy over the physical phenomenon that underlies these \nhelicity -dependent all optical processes . \n 2 \n I. INTRODUCTION \nAll optical switching (AOS) refers to the manipulation of the magnetic state of a \nsample through the application of short laser pulses. The discovery of subpicosecond \ndemagnetization of a nickel sample [1] upon application of a short laser pulse, ranging \nfrom tens of femtosecond to several picoseconds , opened the path for other experiments \nto manipulate the magnetization using ultrashort laser pulses in ferromagnetic [2,3] , \nsynthetic antiferromagnetic [4–6] and ferrimagnetic materials [7 –9]. While the AOS in \nferromagnetic materials is usually described by the Magnetic Circular Dichrois m \n(MCD) [10,11] or the Inverse Faraday Effect (IFE) [12–14] , and it requires multiple \nshots of circularly polarized laser pulses [15] , the inversion of the magnetization of \nferrimagnetic materials can be achieved by single- shot pulse [16] , even with linear \npolarization. In these Helicity -Independent AOS (HI-AOS) processes, the reversal takes \nplace as the two antiferromagnetically coupled sublattices demagnetize at different rates \nwhen submitted to a laser pulse of adequate duration and energy. Since exchange \nprocesses conserve total angular momentum, the system transits through a ferromagnetic -\nlike state despite being ferrimagnetic at the ground state [17] . The switching of the \nmagnetization is completed when the sublattices relax back to their thermodynamic \nequilibrium [18] . On the other hand, several experimental studies [7,9] have also \nobserved that the magnetic state of ferrimagnetic alloys can be also reversed under \ncircular polarized laser pulses within a narrow range of laser energies, resulting in a \nHelicity -Dependent AOS (HD -AOS ) which could be useful to develop ultrafast magnetic \nrecording devices purely controlled by optical means. While the single -shot HI -AOS can \nbe caused by the strong non- equilibrium due to the heating induced by the laser pulse, the \nphysical mechanisms behind the HD -AOS are still not completely understood, and \nseveral works participated by the same authors ascribe it either to the IFE [9] or the \nMCD [8]. Although several attempts have been performed to explain such AOS \nprocesses, a realistic numerical description of experimental observations is still missing . \nIndeed, some theoretical studies usually adopt an atomistic description which is limited \nto small samples [16,19] , with dimensions at the nanoscale, well below the size of the \nexperimentally studie d samples, with lateral sizes of several hundreds of microns . Such \natomistic approach cannot describe the non -uniform heating caused by laser beams of \nseveral microns, so it does not predict some multidomain patterns typically observed in \nthe experiments [9]. On the other hand, other numerical attempts have been carried out 3 \n by describing the ferrimagnetic alloy as an effective ferromagnetic sample, without \nconsidering the individual nature of the two sublattices forming the ferrimagnet [9] . \nAlthough these micromagnetic studies predict some features of the AOS processes, so \nfar, the structure of the ferrimagnetic alloys has not been taken into account to investigate \nthe reversal of magnetic samples of micrometer size under realistic excitation conditions . \nAs the switching happens due to angular momentum transfer between subl attices , \nsomething impossible to account for within an effective ferromagnetic description, it is needed to develop studies considering the two sublattice nature of the such alloys to \nnaturally evaluate their role on the reversal processes. \nHere we present a micromagnetic framework that is able to reproduce accurately \nthe atomistic results of the laser -induced switching by the extension of the conventional \nLandau -Lifshitz -Bloch (LLB) model for ferrimagnets [20,21] . Note that the conventional \nLLB model does not allow to accurately describe AOS as indicated in [21] , and here we \nextend it to solve this limitation . However, and differently from atomistic simulations, \nwhich are limited to small samples at the nanoscale submitted to uniform laser heating, \nour micromagnetic formalism allows us to realistically describe AOS experimental \nobservations by directly evaluating extended samples at the microscale and non- uniform \nenergy absorption from the laser pulse . The procedures here developed are essential to \nunderstand the physical aspects underlying these experiments , and will be useful for the \nfuture development of novel ultra -fast devices based on these AOS processes. After \npresenting and validating both the atomistic and the extended micromagnetic models for \nsample size of ten s of nanometers, the upper size limit of the atomistic spin models , we \ndescribe the results for HI -AOS processes in realisti c samples at the microscale for a \ntypically ferrimagnetic alloy (GdFeCo) . Later on, we focus our attention to the description \nof the HD -AOS processes by exploring the role of the IFE and MCD separately for two \ndifferent ferrimagnetic alloys where the relative composition is slightly varied. Our \nresults allow us to suggest future experiments which could be useful to infer the \ndominance of the IFE or the MCD in single -shot HD-AOS in ferrimagnetic alloys . \n \nII. ATOMISTIC AND MICROMAGNETIC MODELS \nTypical ferrimagnetic (FiM) samples formed by a Transition Metal (TM:Co, CoFe) \nand a Rare Earth (RE:Gd) are considered here. Square samples in the 𝑥𝑥𝑥𝑥 plane with side 4 \n length ℓ and with thickness 𝑡𝑡𝐹𝐹𝐹𝐹𝐹𝐹 =5.6 nm are studied. At atomistic level the FiM sample \nis formed by a set of coupled spins, and the magnetization dynamics is described the \nLangevin -Landau -Lifshitz -Gilbert eq uation \n𝜕𝜕𝑆𝑆⃗𝐹𝐹\n𝜕𝜕𝑡𝑡=−𝛾𝛾0\n(1+𝜆𝜆2)�𝑆𝑆⃗𝐹𝐹×�𝐻𝐻��⃗𝐹𝐹+𝐻𝐻��⃗𝑡𝑡ℎ,𝐹𝐹 �+𝜆𝜆𝑆𝑆⃗𝐹𝐹×�𝑆𝑆⃗𝐹𝐹×�𝐻𝐻��⃗𝐹𝐹+𝐻𝐻��⃗𝑡𝑡ℎ,𝐹𝐹 ��� (1) \nwhere 𝑆𝑆⃗𝐹𝐹 is the localized magnetic moment and 𝐻𝐻��⃗𝐹𝐹 is the local effective field including \nintra- and inter -lattice exchange and anisotropy contributions . 𝐻𝐻��⃗𝑡𝑡ℎ,𝐹𝐹 is the local stochastic \nthermal field . 𝛾𝛾0 and 𝜆𝜆 are the gyromagnetic ratio and the damping parameter \nrespectively [22] . Except the contrary is indicated, t ypical parameters of Gdx(CoFe )1-x \nwith relative composition 𝑥𝑥 =0.25 were considered [22] . See Supplemental Material \nNote SN1 [23] for further details , including material and numerical parameters . \nStarting from an initial uniform state of the FiM with the spins of the two sublattices \nantiparallelly aligned each other along the easy axis 𝑧𝑧, a laser pulse is applied , and the \nirradiated sample absorbs energy from the laser pulse. The laser spot is assumed to have \na spat ial Gaussian profile ( 𝜂𝜂(𝑟𝑟)), with 𝑟𝑟0 being the radius spot (𝑑𝑑0=2𝑟𝑟0 is the full width \nat half maximun, FWHM). Its temporal profile ( 𝜉𝜉(𝑡𝑡)) is also Gaussian, with 𝜏𝜏𝐿𝐿 \nrepresenting the pulse duration (FWHM). The absorbed power density can be expressed \nas 𝑃𝑃(𝑟𝑟,𝑡𝑡)=𝑄𝑄𝜂𝜂(𝑟𝑟)𝜉𝜉(𝑡𝑡) where 𝜂𝜂(𝑟𝑟)=exp[−4ln(2)𝑟𝑟2/ (2𝑟𝑟0)2 ] is the spatial profile \nwith 𝑟𝑟=�𝑥𝑥2+𝑥𝑥2 being the distance from the center of the laser spot, and 𝜉𝜉 (𝑡𝑡)=\nexp[−4ln(2)(𝑡𝑡−𝑡𝑡0)2/ 𝜏𝜏𝐿𝐿2] the temporal profile. 𝑄𝑄 is the maximum value of the \nabsorbed power density reached at 𝑡𝑡=𝑡𝑡0 just below the center of the laser spot. \nLaser pulse heats the FiM sample, and consequently, it is transiently dragged into a \nnon-equilibrium thermodynamic state, where its magnetization changes according to the \ntemperature dynamics. The temperature evolution is described by the Two Temperatures Model (TTM) [9,24] in terms of two subsystems: the electron ( 𝑇𝑇\n𝑒𝑒=𝑇𝑇𝑒𝑒(𝑟𝑟⃗,𝑡𝑡)) and the \nlattice (𝑇𝑇𝑙𝑙=𝑇𝑇𝑙𝑙(𝑟𝑟⃗,𝑡𝑡)), \n𝐶𝐶𝑒𝑒𝜕𝜕𝑇𝑇𝑒𝑒\n𝜕𝜕𝑡𝑡=−𝑘𝑘𝑒𝑒∇2𝑇𝑇𝑒𝑒−𝑔𝑔𝑒𝑒𝑙𝑙(𝑇𝑇𝑒𝑒−𝑇𝑇𝑙𝑙)+𝑃𝑃(𝑟𝑟,𝑡𝑡)−𝐶𝐶𝑒𝑒(𝑇𝑇𝑒𝑒−𝑇𝑇𝑅𝑅)\n𝜏𝜏𝐷𝐷 (2) \n𝐶𝐶𝑙𝑙𝜕𝜕𝑇𝑇𝑙𝑙\n𝜕𝜕𝑡𝑡=−𝑔𝑔𝑒𝑒𝑙𝑙(𝑇𝑇𝑙𝑙−𝑇𝑇𝑒𝑒) (3) \nwhere 𝐶𝐶𝑒𝑒 and 𝐶𝐶𝑙𝑙 denote the specific heat of electrons and lattice subsystems, respectively. \n𝑘𝑘𝑒𝑒 is the electronic thermal conductivit y. 𝑔𝑔𝑒𝑒𝑙𝑙 is a coupling parameter between the electron 5 \n and lattice subsystems, and 𝜏𝜏𝐷𝐷 is the characteristic heat diffusion time to the substrate and \nthe surrounding medi a [25] . Conventional values were adopted (see [1,16,26] and \nSupplemental Material Note SN1 [23]). \nThe approach that consists on solving Eq. (1 ) coupled to Eqs. ( 2)-(3) is named as \nAtomistic Spin Dynamics (ASD), and due to computational restrictions, its numerical \nsolution is limited to small samples at the nanoscale ( ℓ ≲ 100 nm, see Supplemental \nMaterial Note SN2 [23]) . While A SD predicts the single -shot switching in small FiM \nnano- samples [16,22,27] , the lack of a realistic micromagnetic framework for micro -size \nsamples and non- unifom laser spot limits the description of many experimental \nworks [9] . In particular, the appearance of central regions with a multi -domain \ndemagnetized pattern s [28] , or the observation of rings of switched magnetization under \nirradiation with lasers of tens of micrometers [29] cannot be reproduced by AS D due to \nsuch computing limitations. In order to overcome the ASD limitations , here we develop \nan extended continuous micromagnetic model that describes the temporal evolution of \nthe reduced local magnetization 𝑚𝑚��⃗𝐹𝐹(𝑟𝑟 ⃗,𝑡𝑡) of each sublattice 𝑖𝑖 :RE,TM based on the \nconventional ferrimagnetic Landau -Lifshitz -Bloch (LLB) Eq [21,30], \n𝜕𝜕𝑚𝑚��⃗𝐹𝐹\n𝜕𝜕𝑡𝑡=−𝛾𝛾0𝐹𝐹′�𝑚𝑚��⃗𝐹𝐹×𝐻𝐻��⃗𝐹𝐹�+ \n−𝛾𝛾0𝐹𝐹′𝛼𝛼𝐹𝐹⊥\n𝑚𝑚𝐹𝐹2𝑚𝑚��⃗𝐹𝐹×�𝑚𝑚��⃗𝐹𝐹×�𝐻𝐻��⃗𝐹𝐹+𝜉𝜉⃗𝐹𝐹⊥��+ \n+𝛾𝛾0𝐹𝐹′𝛼𝛼𝐹𝐹∥\n𝑚𝑚𝐹𝐹2(𝑚𝑚��⃗𝐹𝐹·𝐻𝐻��⃗𝐹𝐹)𝑚𝑚��⃗𝐹𝐹+𝜉𝜉⃗𝐹𝐹∥ (4) \nwhere 𝐻𝐻��⃗𝐹𝐹=𝐻𝐻��⃗𝐹𝐹(𝑟𝑟 ⃗,𝑡𝑡) is the local effective field on sublattice magnetic moment 𝑖𝑖 at location \n𝑟𝑟 ⃗ of the FiM sample , 𝛼𝛼𝐹𝐹∥ and 𝛼𝛼𝐹𝐹⊥ are the longitudinal and perpendicular damping \nparameters, and 𝜉𝜉⃗𝐹𝐹∥ and 𝜉𝜉⃗𝐹𝐹⊥ are the longitudinal and perpendicular stochastic thermal fields. \nDetails of the LLB model can be found in [21,30] and in Supplemental Material Note \nSN2 [23]. C ontrary to the ASD, where the spatial discretization is imposed by the \natomistic scale ( 𝑎𝑎= 0. 35 nm ), within the micromagnetic model the sample is discretized \nin elementary cells with dimensions of Δ𝑥𝑥=Δ𝑥𝑥~1nm a\nnd Δ𝑧𝑧=𝑡𝑡𝐹𝐹𝐹𝐹𝐹𝐹. Therefore, it is \npossible to numerically evaluate, with manageable computing effort, extended samples at \nthe microscale (ℓ~100 μm), three order s of magnitude larger than the ones which can be \ndealt with the ASD model. 6 \n Numerically solving Eq. (4) coupled to Eqs . (2)- (3) under ultra -short laser pulses \nprovides a micromagnetic description of several AOS processes in ferromagnetic \nsystems [13] . However, when dealing with ferr imagnetic samples we checked that some \ndisagreement with the predictions of the ASD mode l were observed (see Supplemental \nMaterial Note SN3 [23] ), which are related to the lack of a proper description of the \nangular moment exchange between sublattices during the non- equilibrium transient state \npromoted by the laser pulse. Indeed, magnetization dynamics in FiMs is driven by \ndissipative processes of relativistic and exchange nature. The relativistic ones allow \nexchange of angular momentum between the magnetization and the lattice degree of \nfreedom due to the spin- orbit coupling between them, and are phenomenologically \ndescribed by the usual damping terms in the LLB Eq. (4). Additiona lly, in multisublattice \nmagnets as FiMs, another different pathway opens local exchange of angular momentum \nbetween both sublattices of the FiM, and to account for it, the LLB Eq . (4) has to be \nenhanced by an additional exchange relaxation torque [18,31–34] . The simplest model \nto describe the sublattice -specific magnetization dynamics in FiMs, was derived from \nOnsager’s relations [31] within a macrospin approach based on a microscopic spin \nmodel. In this simplified description, the magnetization dynamics of sublattice 𝑖𝑖 can be \nexpressed as 1\n𝛾𝛾0𝑖𝑖𝑑𝑑𝑚𝑚𝑖𝑖\n𝑑𝑑𝑡𝑡=𝛼𝛼𝐹𝐹𝐻𝐻𝐹𝐹+𝛼𝛼𝑒𝑒𝑒𝑒�𝜇𝜇𝑖𝑖\n𝜇𝜇𝑗𝑗𝐻𝐻𝐹𝐹−𝐻𝐻𝑗𝑗�, where 𝑖𝑖,𝑗𝑗:RE,TM, 𝜇𝜇𝐹𝐹 and 𝐻𝐻𝐹𝐹 are the \nmagnitude of the magnetic moment and the e ffective field acting on macrospin of \nsublattice 𝑖𝑖 respectively . The relativistic relaxation parameter in this model, 𝛼𝛼𝐹𝐹, \ncorresponds to the longitudinal damping parameter in the LLB E q. and it depends on the \ntemperature of the thermal bath to which ang ular momentum and energy is dissipated. \nDifferently, it is assumed that the exchange relaxation parameter 𝛼𝛼 𝑒𝑒𝑒𝑒, only depends on \nthe non- equilibrium sublattice magnetizations, 𝛼𝛼𝑒𝑒𝑒𝑒=𝛼𝛼𝑒𝑒𝑒𝑒(𝑚𝑚𝐹𝐹,𝑚𝑚𝑗𝑗). Considering that \nexchange relation rate should be symmetric with respect to the sublattice index, \n𝛼𝛼𝑒𝑒𝑒𝑒(𝑚𝑚𝐹𝐹,𝑚𝑚𝑗𝑗)=𝛼𝛼𝑒𝑒𝑒𝑒(𝑚𝑚𝑗𝑗,𝑚𝑚𝐹𝐹), a simple functional fulfilling these heuristic conditions \nyields to 𝛼𝛼𝑒𝑒𝑒𝑒(𝑚𝑚𝐹𝐹,𝑚𝑚𝑗𝑗)=𝜆𝜆𝑒𝑒𝑒𝑒𝑚𝑚𝑖𝑖+(𝑒𝑒𝑗𝑗𝜇𝜇𝑗𝑗/𝑒𝑒𝑖𝑖𝜇𝜇𝑖𝑖 )𝑚𝑚𝑗𝑗\n𝑚𝑚𝑖𝑖𝑚𝑚𝑗𝑗 where 𝜆𝜆𝑒𝑒𝑒𝑒 is a phenomenological parameter \nrepresenting the exchange relaxation rate and 𝑥𝑥𝐹𝐹 the concentration of each specimen 𝑖𝑖. \nInspired by this two sublattice phenome nological model based on Onsager’s relations, \nhere we add an additional torque 𝜏𝜏⃗𝐹𝐹𝑁𝑁𝑁𝑁 to the micromagnetic LLB Eq. (4) that accounts for \nnon-equilibrium magnetic moment exchange between sublattices , and becomes crucial to \ndescribe AOS ultra -fast switching in FiMs under realistic conditions . The torque reads as 7 \n 𝜏𝜏⃗𝐹𝐹𝑁𝑁𝑁𝑁=𝛾𝛾0𝐹𝐹′𝜆𝜆𝑒𝑒𝑒𝑒𝛼𝛼𝐹𝐹∥𝑥𝑥𝐹𝐹𝜇𝜇𝐹𝐹𝑚𝑚𝐹𝐹+𝑥𝑥𝑗𝑗𝜇𝜇𝑗𝑗𝑚𝑚𝑗𝑗\n𝜇𝜇𝐹𝐹𝑚𝑚𝐹𝐹𝜇𝜇𝑗𝑗𝑚𝑚𝑗𝑗�𝜇𝜇𝐹𝐹𝐻𝐻��⃗𝐹𝐹∥−𝜇𝜇𝑗𝑗𝐻𝐻��⃗𝑗𝑗∥� (5) \nwhere and 𝐻𝐻��⃗𝐹𝐹∥ and 𝐻𝐻��⃗𝑗𝑗∥ are the longitudinal effective field s for each lattice 𝑖𝑖:RE,TM [21] , \n𝑥𝑥𝐹𝐹≡𝑥𝑥 and 𝑥𝑥𝑗𝑗=1−𝑥𝑥𝐹𝐹=1−𝑥𝑥 are the concentration s of each specimen, and 𝜆𝜆𝑒𝑒𝑒𝑒 is a \nparameter representing the exchange relaxation rate [18] . By including Eq. (5) in the \nRHS of Eq. (4), and numerically solving it coupled to TTM Eqs. (2) -(3), w e can provide \na realistic description of the magnetization dynamics in FiM systems under ultra -short \nlaser pulses. In what follows, we refer to this formalism as the extended micromagnetic \nLLB model (eLLB) , to distinguish it from the conventional LLB model (LLB) when \n𝜏𝜏⃗𝐹𝐹𝑁𝑁𝑁𝑁=0. See Supplemental Material Note SN2 [23] for the rest of details. \n \nIII. RESULTS AND DISCUSSION \nBefore presenting the predictions of the extended micromagnetic model for realistic \nFiM samples and laser beam s at the microscale, here we firstly compare the results \nobtained from the extended LLB model ( eLLB) to the ones resulting from the atomistic \nspin dynamics simulations (ASD) for a small FiM dot at the nano -scale (ℓ≈25 nm ). As \ntypical laser spots have radius of 𝑟𝑟0~1 μm−10 μm or even larger , we assume here that \nthe power absorbed by the FiM dot the from the laser pulse is uniform, that is, 𝜂𝜂(𝑟𝑟)=1. \nThe pulse duration is 𝜏𝜏 𝐿𝐿=50 fs. Typical results showing the temporal evolution of the \nout-of-plane averaged mag netization ( 𝑚𝑚𝑧𝑧𝐹𝐹) for each sublattice ( 𝑖𝑖:TM (red), RE (blue)) are \nshown in FIG. 1 for two different values of 𝑄𝑄. 8 \n \nFIG. 1 . Comparison between atomistic simulation (ASD, solid lines) and micromagnetic model \n(eLLB , dashed lines) results for the temporal evolution of the out -of-plane magnetization for the \ntransition metal (TM, red) and the rare earth (RE, blue) under two different laser power densities: \n(a) 𝑄𝑄=6×1021 W/m3 and (b) 𝑄𝑄=12×1021 W/m3. Four consecutive laser pulses with 𝜏𝜏𝐿𝐿=\n50 fs are applied every 20 ps. (c) shows a detailed view of first pulse switching event as in ( a), \nwhile (d) shows temporal evolution of the electron ( 𝑇𝑇𝑒𝑒) and lattice ( 𝑇𝑇𝑙𝑙) temperatures for 𝑄𝑄=\n6×1021 W/m3. (e) and (f) corresponds to 𝑄𝑄=12×1021 W/m3. Shaded interval in (e) shows \nthe transient ferromagnetic state. The pulse length is 𝜏𝜏𝐿𝐿=50 fs. The eLLB results were obtained \nwith 𝜆𝜆𝑒𝑒𝑒𝑒=0.013. \n \nA remarkable agreement between both ASD and e LLB models with similar \ndynamics for both sublattices is observed in FIG. 1 (a) and (b) . For low 𝑄𝑄 values ( FIG. \n1(a) and (c) ) there is no switching, but when 𝑄𝑄 increases above a threshold, which \ndepends on the pulse length ( 𝜏𝜏𝐿𝐿), the deterministic AOS is predicted by both ASD and \neLLB models ( FIG.1(b) and (e) ). It is important to note that similar switching was also \nobtained within the deterministic e LLB framework, that is, in the absence of thermal \n9 \n fluctuations (𝜉𝜉⃗𝐹𝐹⊥=𝜉𝜉⃗𝐹𝐹∥=0 in Eq. (4) , see FIG. S4(b) in Supplemental Material Note \nSN3 [23] ). On the contrary, the conventional LLB model ( LLB, 𝜏𝜏⃗𝐹𝐹𝑁𝑁𝑁𝑁=0) fails to \nreproduce the switching of FIG. 1(b) (see FIG. S4(c) -(d) in Supplemental Note \nSN3 [23] ). FIG. 1(c) show s the details of the temporal evolution 𝑚𝑚𝑧𝑧𝐹𝐹 for 𝑄𝑄=\n6×1021 W/m3 during the first laser pulse , while the corresponding evolutions of 𝑇𝑇𝑒𝑒 and \n𝑇𝑇𝑙𝑙 are depicted in FIG. 1(d) , which also shows the laser pulse . Corresponding results for \n𝑄𝑄=12×1021 W/m3 are shown in FIG. 1(e) and (f) respectively . For 𝑄𝑄=\n6×1021 W/m3, the electron temperature reaches a peak maximum value of 𝑇𝑇𝑒𝑒≈850 K \nat the end of the laser pulse, but this is not enough to achieve the switching . For 𝑄𝑄=\n12×1021 W/m3, 𝑇𝑇𝑒𝑒 reaches a peak of 𝑇𝑇𝑒𝑒≈1150 K and switching takes place . Not ice \nthat this value is well above the Curie temperature ( 𝑇𝑇𝐶𝐶≈600 K), and therefore the system \nneeds to be significantly heated above the Curie threshold to achieve the deterministic \nAOS in FiM . These processes are explained by the different demagnetization rates of the \nRE and TM sublattices, that lead to a transient ferromagnetic aligment. Such transient \nferromagnetic state is observed during a short transient (see shaded interval in FIG. 1(e)), \nand it is only present when the system is far away from the thermodynamic equilibrium, \nas caused by the ultrafast laser heating. Except the contrary is indicated, all eLLB results \nwere obtained wi th 𝜆𝜆𝑒𝑒𝑒𝑒=0.013, (s ee FIG. S 5 and its corresponding discussion in \nSupplemental Material Note SN3 (c) [23] for results with other values of 𝜆𝜆𝑒𝑒𝑒𝑒). \n \nHelicity -Independent A ll Optical Switching (H I-AOS) . Once validated the e LLB \nformalism by reproducing the ASD results for small nano- sample s under uniform \nlinearly -polarized laser pulses, we can now use it to explore the influence of the laser \nduration ( 𝜏𝜏𝐿𝐿) and maximum absorbed power density (𝑄𝑄) in realistic extended samples at \nthe microscale (ℓ ~10 μ m). This is illustrated in the phase diagram of FIG. 2(a), which \nshows the final state under a single linearly -polarized laser pulse starting from a uniform \nstate of the FiM . White color indicates the combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) where the sample \nreturns to the original state after the pulse (no-switching) . The blue region corresponds to \ncombinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) presenting deterministic HI-AOS after each pulse, and red \ncorresponds to combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) resulting in a final demagnetized multidomain \nconfiguration. It is noted that there is a correlation b etween the final state and the \nmaximum electron temperature reached in the sample, which is shown by the overlapping 10 \n solid black lines in FIG. 2(a). As it is clearly observed, s olid lines coincide with \nboundaries between the three possible behaviors already discussed. Indeed, the transition \nbetween no -switching (white) to the deterministic switching range (red) is limited by the \n~1000 K curve , whereas the transition to the thermal demagnetization (blue) occurs when \n𝑇𝑇𝑒𝑒≳1400 K, as shown in FIG. 2(a). Instead of 𝑄𝑄 , the information collected in the phase \ndiagram of FIG. 2(a) could be also presented in terms of the laser fluence ( 𝐹𝐹≡𝑄𝑄 𝜏𝜏𝐿𝐿 𝑡𝑡𝐹𝐹𝐹𝐹𝐹𝐹), \nas it is done in FIG. S6 of Supplemental Material Note SN 4 [23] . Note, that such phase \ndiagram is also in good qualitative agreement with recent experimental \nobservations [35] . \n \nFIG. 2 . (a) Phase diagram of the final state as a function of 𝑄𝑄 and 𝜏𝜏𝐿𝐿 for a small nano -sample ℓ=\n25 nm under uniform laser heating ( 𝜂𝜂(𝑟𝑟)=1). White, red and blue colors represent no -\nswitching, deterministic switching and thermal demagnetization behaviors respectively. Solid \nlines are isothermal curves showing the maximum electron temperature ( 𝑇𝑇𝑒𝑒) reached due to the \npulse. (b) Typical mi cromagnetic snapshots of the initial and final magnetization of RE and TM \nfor three combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿). I: (8×1021 W/m3,20 fs), II: (15×1021 W/m3,30 fs) and \nIII: (20×1021 W/m3,60 fs). Here extended samples ( ℓ=20 μm) with a laser spot of 𝑑𝑑0=ℓ/2 \nwere considered. Dashed purple lines in the images of the initial state indicate the FWHM of the \nlaser spot . The results of the phase diagram (a) coincide with (b) for magnetization at the center \nof the laser spot . \n \nThe main advantage of the extended e LLB model over ASD simulations is that it \nallows us to explore realistic samples and laser beams with dimensions that are not \naccessible with ASD models. e LLB model ( Eqs. (4) and (2) -(3)) has been used to simulate \n11 \n samples with lateral size of ℓ=20 μ m. From no w on , the spatial Gaussian dependence \nof the laser beam is considered (𝜂𝜂(𝑟𝑟)=exp[−4ln(2)𝑟𝑟2/ (2𝑟𝑟0)2 ]), with a laser spot \ndiameter of 𝑑𝑑0=2𝑟𝑟0=ℓ/2. Typical initial and final states corresponding to three \nrepresentative combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) are shown in FIG. 2(b). Our micromagnetic \nsimulations point out again that the three types of behaviors observed experimentally (see \nfor example Fig. 4(a) in [9] or [6,28] ) are also achieved under these realistic conditions, \nwith samples and laser spots at the microscale. Not ice that now the final magnetic state \ndepends on the local position because the power absorption from laser pulse does . The \nfinal states depict a radial symmetry around the center of the laser spot, which coincides \nwith the center of the FiM sample at (𝑥𝑥𝑐𝑐,𝑥𝑥𝑐𝑐)=(0,0). \nIn order to further describe such spatial dependence, FIG. 3 plots the final state of \n𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹 as a function of 𝑥𝑥 along the central line of the FiM sample ( 𝑥𝑥=0) for the same \nthree representative combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) as in FIG. 2(b). The maximum electron \ntemperature 𝑇𝑇𝑒𝑒=𝑇𝑇𝑒𝑒(𝑥𝑥) is also plotted in top graphs by blue curves. Bottom graphs in \nFIG. 3 show the final state over the sample plane (𝑥𝑥,𝑥𝑥). These graphs clearly correlate \nthe local final magnetic state ( 𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹(𝑥𝑥,𝑥𝑥)) with the maximum electron temperature \n𝑇𝑇𝑒𝑒(𝑥𝑥,𝑥𝑥). In the no- switching regime (I), the electron temperature does not reach 1000 K \nat any point . For combinations (𝑄𝑄,𝜏𝜏𝐿𝐿) as II, 𝑇𝑇𝑒𝑒(𝑥𝑥)≳1000 K is only reached in the cent ral \nregion , whose dimensions fit the local part of the sample that switches its magnetization. \nNote that 𝑇𝑇 𝑒𝑒 remains below 𝑇𝑇 𝑒𝑒(𝑥𝑥)≲1400 K. Finally, the demagnetiz ed case (multi -\ndomain pattern, III) occurs in the part of the sample where 𝑇𝑇𝑒𝑒≳1400 K, but deterministic \nswitching is s till obtained in the ring region, where 1000 K≲𝑇𝑇𝑒𝑒(𝑥𝑥)≲1400 K. \nMicromagnetic images (bottom graphs in FIG. 3 ) are in good agreement with typical \nexperimental HI -AOS observations [6,28] . 12 \n \nFIG. 3 . Final out -of-plane magnetization ( 𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹(𝑥𝑥)) and maximum electron temperature ( 𝑇𝑇𝑒𝑒(𝑥𝑥)) \nas function of 𝑥𝑥 for 𝑥𝑥=0 and for the t hree representative combinations of (𝑄𝑄,𝜏𝜏𝐿𝐿) as in FIG. 2(b) \n(top graphs). The corresponding final states over the sample plane (𝑥𝑥,𝑥𝑥) are shown in bottom \ngraphs. \n \nThe inferred correlation between the maximum electron temperature and the final \nmagnetic state allows us to predict the size of the inverted region by studying the \nmaximum electron temperature reached in the sample by using the TTM (Eqs. (2)-(3)) in \ncombination with t he switching diagram of FIG. 2(a) . The radius of the switched area \n(𝑅𝑅𝑠𝑠) calculated from micromagnetic simulations (dots, e LLB), and the one predicted by \nthe TTM (lines, TTM) is shown in FIG. 4 as function of 𝜏𝜏 𝐿𝐿 for two different values of 𝑄𝑄 \nwithin the deterministic switching range (1000 K≲𝑇𝑇𝑒𝑒(𝑥𝑥)≲1400 K). Again, good \nagreement is obtained , a fact that points out that the origin of these HI -AOS processes \nunder linearly polarized laser pulse is a purely thermal phenomenon. Indeed, as the local \nmaximum electron temperature reached in the sample only depends on the absorbed \npower from the laser ( 𝑄𝑄) and the laser pulse length ( 𝜏𝜏𝐿𝐿), the size of the inverted region \ncan be directly obtained from the TTM by the condition of 𝑇𝑇 𝑒𝑒≳1000 K. \n13 \n \nFIG. 4 . Radius of the switched area ( 𝑅𝑅𝑠𝑠) as a function of laser duration ( 𝜏𝜏𝐿𝐿) for two values of the \nmaximum absorbed power ( 𝑄𝑄). Dots are micromagnetic results from the e xtended e LLB model . \nLines are predictions from the TTM, where 𝑅𝑅𝑠𝑠 is inferred from the condition that the local \nmaximum electron temperature reaches 𝑇𝑇𝑒𝑒≳1000 K. \n \nHelicity -Dependent All Optical Switching (H D-AOS) . Previous results were \ncarried out by applying laser pulses with linear polarization (𝜎𝜎=0), and show that the \nHI-AOS can be achieved in a controlled manner with an adequate election of the laser \npower (𝑄𝑄) and duration ( 𝜏𝜏𝐿𝐿): the magnetization switches its direction in the p icoseconds \nrange independently on the initial state. While this is interesting for toogle memory \napplications, the procedure to store and record a bit using linearly polarized laser pulses \nwould still require two steps: ( 𝑖𝑖) a pre liminary reading operation of the magnetic state, \nand after that, ( 𝑖𝑖𝑖𝑖) deciding or not to apply the laser pulse depending on the preceeding \nstate. This two -step procedure can be avoided by using circularly polarized laser pulses, \nresulting in Helicity -Dependent AOS processes (HD- AOS) . However, as it was already \ncommented the physics behind these HD -AOS observations still remains un clear, and \nboth the Magnetic Circular Dichroism (MCD) [8,10] and the Inverse Faraday Effect \n(IFE) [12,14] have been suggested as responsible of the experimental observations. In \nwhat follows, we explore both mechanisms in a separated manner by including them in \nthe extended micromagnetic model. \nLet firstly consider the Magnetic Circular Dichroism . It has been suggested to be \nplay a dominant a role on these HD- AOS processes in GdFeCo ferrimagnetic samples, \nwhich are known for its strong magneto- optical effect [8]. According to the MDC \nformalism, right -handed (𝜎𝜎+) and left-handed (𝜎𝜎−) circular ly polarized laser pulses \n14 \n experience different refractive indices, and consequently a differen ce in energy \nabsorption of the FiM sample for 𝜎𝜎+ and 𝜎𝜎− pulses is expected. The MCD coefficient \ncan be calculated from the total absorption for each polarization, resulting in MCD≡𝑘𝑘=\n(𝐴𝐴−−𝐴𝐴+)/(1\n2(𝐴𝐴++𝐴𝐴−)), where 𝐴𝐴± represent the total absorption for each polarization, \n(±≡𝜎𝜎±). Indeed, the MCD makes the power absorbed by the sample ( 𝑃𝑃(𝑟𝑟,𝑡𝑡)) to depend \non the laser helicity ( 𝜎𝜎±=±1 for right -handed and left -handed circular helicities ) and \non the initial net magnetic state ( 𝑚𝑚𝑁𝑁(0)=𝑀𝑀𝑆𝑆𝑇𝑇𝐹𝐹𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹(0)+𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁(0)), up (↑: \n𝑚𝑚𝑁𝑁(0)>0) or down (↓:𝑚𝑚𝑁𝑁(0)<0). Note that 𝑚𝑚 𝑧𝑧𝑇𝑇𝐹𝐹(0)=±1 and 𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁(0)=∓1, \nwhereas 𝑀𝑀𝑆𝑆𝑇𝑇𝐹𝐹 and 𝑀𝑀𝑆𝑆𝑅𝑅𝑁𝑁 are both positive. Under a right -handed laser pulse (𝜎𝜎+), an \ninitially up (down) magnetic state is expected to absorb more (less) energy than the \ninitially down (up) state . Therefore, 𝑃𝑃(𝑟𝑟,𝑡𝑡) in Eq. (2) is replaced by 𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁)𝑃𝑃(𝑟𝑟,𝑡𝑡) \nwith 𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁)=�1+12𝑘𝑘𝜎𝜎±sign (𝑚𝑚𝑁𝑁)� describing the different absorption power for \nup and down magnetization states as depending on the laser helicity. See further details \non the implementation of the MCD in Supplemental Material Note SN 5 [23] . \nWe have evaluated the role of the MCD in the eLLB model with several values of \nthe MCD coefficient (𝑘𝑘). The isothermal curve delimiting the border between the no-\nswitching and switching regimes now depends on the combination of helicity and initial \nnet magnetic state (see such isothermal threshold curves for different values of the MCD \ncoefficient in of SN 5). Considering a realistic value of MCD≡𝑘𝑘~2%, as estimated \nin [8], the electron temperature variation is quite small, typically a few units of K, and \ntherefore small variations in the phase diagram are obtained with respect to the one for \nlinearly -polarized pulses FIG. 2(a) (see also FIG. S 7 in Supplemental Material Note \nSN5 [23] ). However, when exciting with circular polarized pulses close to the no-\nswitching /switching boundary, the FiM switches or not depending on the helicity and \ninitial net state, only within a narrow interval of 𝑄𝑄. This is represented in FIG. 5(a), where \nthe H D-AOS is shown for 𝜏𝜏 𝐿𝐿=50 fs pulses with different 𝑄𝑄 in a sample with ℓ=20 μm. \nNote that these results correspond to a F iM alloy Gdx(FeCo) 1-x with 𝑥𝑥=0.25, and that for \nthis relative composition the RE is the dominant sublattice at room temperature, 𝑀𝑀 𝑠𝑠𝑅𝑅𝑁𝑁>\n𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹 at 𝑇𝑇=300 K (see inset of FIG. S3 of Supplemental Material [23] or FIG. 6(a) ). No \nswitching is achieved for low energy values (see left colum in FIG. 5(a) for 𝑄𝑄=\n5.7×1021 W/m3). However, if 𝑄𝑄 increases to 𝑄𝑄 =5.8×1021 W/m3, the system shows \nthe so -called H D-AOS: if 𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁 is initially down (up), the reversal is only achieved for 15 \n left-handed helicity, 𝜎𝜎−=−1 (right -handed helicity, 𝜎𝜎+=+1). Consequently, the final \nstate can be selected by chosing the laser helicity, which is relevant for ultrafast memory \napplications. It is important to note that this H elicity Dependent AOS is only obtained in \na very narrow range of 𝑄𝑄 around the H elicity Independent AOS boundary . Indeed, a small \nincrease of the absorbed power results again in H I-AOS as the linear polarized case (see \n3rd and 4th columns in FIG. 5(a) for 𝑄𝑄 =5.9×1021 W/m3 and 𝑄𝑄=9.0×1021 W/m3). \nFor high values of 𝑄𝑄, the final state depicts a ring around a central demagnetized state, \nsimilar to HI -AOS case (see right column in FIG. 5(a) for 𝑄𝑄=18×1021 W/m3). In this \ncase, the maximum electron temperature overcomes the 𝑇𝑇𝑒𝑒≃1400 K threshold in the \ncentral region below the laser spot , resulting in a central demagnetized or multidomain \nstate. However, the maximum 𝑇𝑇𝑒𝑒 remains with in the range of HD -AOS deterministic \nswitching ( 1000 K≲𝑇𝑇𝑒𝑒(𝑥𝑥)≲1400 K) in the ring around the central part . These \nmicromagnetic predictions , including the narrow range of HD -AOS, are in good \nagreement with several experimental observations (see, for instance, Fig. 4 and 7(a) \nin [9] , Fig. 1(b) in [4] , or Fig. 3 in [36] ). \n \n16 \n FIG. 5. Helicity -Dependent AOS predicted by the MCD . (a) Snapshots of the final RE magnetic \nstate (𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁) after a laser pulse of 𝜏𝜏𝐿𝐿=50 fs for five different values of the absorbed power density \n(𝑄𝑄). Results are shown for four combinations of the initial state ( ↑,↓) and helicities ( 𝜎𝜎±) as \nindicated at the left side. The HD- AOS is shown in panel corresponding to 𝑄𝑄=\n5.8×1021 W/m3. (b) RE magnetic state after every pulse for 𝑄𝑄=5.9×1021 W/m3, showing \nthe appearance of a ring due to the MCD and the switching of the central part. Here pulses with \nleft-handed chirality are applied ( 𝜎𝜎+). The sample side is ℓ=20 μm and the laser spot diameter \nis 𝑑𝑑0=ℓ/2. \n \nMoreover, the inclusion of the MCD in our e LLB model allows us to explain the \nexperimental observation of rings [28,36,37] which appear after the application of a \nsecond laser pulse. This is illustrated in FIG. 5(b) for pulses with 𝑄𝑄=5.9×1021 W/m3 \nand 𝜏𝜏𝐿𝐿=50 fs. The central part of the sample reaches temperatures that lead to H I-AOS, \nand ther efore, its magnetization reverses after each pulse. On the contrary, the ring around \nof the inverted region is within the H D-AOS regime , and therefore, its local magnetic \nstate (going from up to down) is only reversed by the first pulse. For the second and \nsubsequent pulses , the ring maintains its down state while the inner part changes again to \nup (white). This is repeated every pulse, with the inversion of the central part and the \nmaintenance of bla ck ring in the external shell, as it is clearly seen in even pulses (see \nsnapshots after pulses #2 and #4 in FIG. 5(b) ). Note that this ring structure differs from \nthe ones shown in FIG. 2(b) and FIG.3 , as they were caused by the inversion of the \nmagnetiza tion around the central demagnetized part under high -power linear pulses (𝜎𝜎=\n0). For circularly polarized pulses (𝜎𝜎±=±1) the images correspond to alternative \nswitching and the H D-AOS without the central demagnetized (multidomain) state . Again, \nthese results are in good agreement with recent experimental observations (see figures \nin [28,36,37] ). \n \nInstead of the MCD, several other works claim that the observations of the HD -\nAOS can be ascribed to the Inverse Faraday Effect (IFE) [9]. Within this formalism, the \nlaser pulse generates an effective out -of-plane magneto -optical field which direction \ndepends on the laser pulse helicity , 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)=𝜎𝜎±𝐵𝐵𝐹𝐹𝑀𝑀𝜂𝜂(𝑟𝑟)𝜖𝜖(𝑡𝑡)𝑢𝑢�⃗𝑧𝑧, where 𝜂𝜂(𝑟𝑟)=\nexp[−4ln(2)𝑟𝑟2/ (2𝑟𝑟0)2 ] is the spatial field profile, and 𝜖𝜖(𝑡𝑡) is its temporal profile. \nNote that the spatial dependence of 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) is the same as the one of the absorbed power \ndensity. However, a ccording to the literature [9], the so- called magneto -optical field \n𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) has some temporal persistence with respect to the laser pulse, and therefore its 17 \n temporal profile is different for 𝑡𝑡 <𝑡𝑡0 and 𝑡𝑡 >𝑡𝑡0: 𝜖𝜖(𝑡𝑡<𝑡𝑡0)=exp[−4ln(2)(𝑡𝑡−𝑡𝑡0)2/\n 𝜏𝜏𝐿𝐿2], and 𝜖𝜖 (𝑡𝑡≥𝑡𝑡0)=exp[−4ln(2)(𝑡𝑡−𝑡𝑡0)2/(𝜏𝜏𝐿𝐿+𝜏𝜏𝐷𝐷)2], where 𝜏𝜏 𝐷𝐷 is the delay time of \nthe 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) with respect to the laser pulse. We have evaluated this IFE scenario by \nincluding this field 𝐵𝐵 �⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) in the effective field of Eq. (4). The results for the same \nFiM alloy considered up to here ( Gdx(FeCo) 1-x, with x=0.25, see SN 5), are similar to the \nones already presented in FIG. 5 for the MCD considering a maximum magneto -optical \nfield of 𝐵𝐵𝐹𝐹𝑀𝑀=20 T with a delay time of 𝜏𝜏𝐷𝐷=𝜏𝜏𝐿𝐿. These IFE results can be seen in FIG. \nS8 in Supplemental Material Note SN6 [23] . Therefore, we could conclude from this \nanalysis that, from the micromagnetic modeling point of view, both the MCD and the IFE \nare compatible with experimental observations of the HD -AOS . At this point, it is worth \nto mention here that in real experiments there is not a clear distinction between MCD and \nIFE phenomena. Indeed, the modeling of the IFE for absorbing materials can account for \nabsorption phenomena as MCD (see for instance [38,39] ). These works suggested that \nin micromagnetic simulations the IFE could induce a change of the magnetic moment \n(Δ𝑚𝑚��⃗𝐹𝐹) modifying the initial magnetic moments in the two sublattices of the FiM when \nsubmitted to circular polarized laser pulses. We have evaluated in our modeling this alternative manner of studying the role of the IFE by adding such an induced magnetic \nmoment in the eLLB Eq. (4), and compared the results to the case where the IFE is \nsimulated by the magneto -optical field 𝐵𝐵\n�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) as discussed above. As presented and \ndiscussed in Supplemental Material Note S N7 [23] , both alternatives ( 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) or Δ𝑚𝑚��⃗𝐹𝐹) \nare equivalent from the simulation point of view. Therefore, in what follows we will \nsimulate the IFE as an effective out -of-plane magneto- optical field. \nIn order to get a further understanding on the physics of these two mechanisms, \neither the MCD or th e IFE, we have explored the switching probability as a function of \n𝑄𝑄 for laser pulses with fixed duration ( 𝜏𝜏𝐿𝐿=50 fs ) in two FiM alloys with slightly \ndifferent composition: Gd x(FeCo) 1-x with x=0.25 and x=0.24 respectively . The \ncorresponding parameters to numerically evaluate these two alloys are given in SN8, and \nthe temperature dependence of the saturation magnetization of each sublattice (RE: Gd; \nTM: CoFe) are shown in FIG. 6 (a) and (c) respectively . Note that magnetization \ncompensation temperature at which the net magnetization of the sample vanishes ( 𝑇𝑇𝐹𝐹) is \nabove and below room temperature for x=0.25 and x=0.24 respectively. In other words, \nthe FiM sample is dominated by the RE (TM) at 𝑇𝑇 =300 K for x=0.25 (x=0.24) \ncompositions . The MCD and IFE parameters remain fixed as indicated above (MCD: 18 \n 𝑘𝑘~2%,; IFE: 𝐵𝐵 𝐹𝐹𝑀𝑀=20 T, 𝜏𝜏𝐷𝐷=𝜏𝜏𝐿𝐿). The two possible initial states prior the laser pulse \n(𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹:(↑,↓)), and the three laser polarizations (linear: 𝜎𝜎 =0, and circular 𝜎𝜎±=±1) were \nevaluated. The switching probability was computed by evaluating ten different stochastic \nrealization s for each for each 𝑄𝑄, and t he results are presented in FIG. 6(b) and (d) for \nx=0.25 and x =0.24 respectively . As in experimental observations [9], the HD -AOS takes \nplace o nly in a narrow range of 𝑄𝑄 around the threshold value of 𝑄𝑄 at which the switching \nprobability abruptly changes from 0 to 1 under linear polarized pulses ( black dots in FIG. \n6(b) and (d)) . \n \n \nFIG. 6. Temperature dependence of the spontaneous magnetization of each sublattice (RE:Gd; \nTM:CoFe) of the FiM alloy (Gd x(CoFe) 1-x) for two different compositions: (a) x=0.25 and (c) \nx=0.24. The vertical grey line indicates the initial room temperature prior the laser pulse ( 𝑇𝑇=\n300 K). Probability of switching as a function of the absorbed power density ( 𝑄𝑄) for a laser pulse \nof 𝜏𝜏𝐿𝐿=50 fs for different combinations of the initial state (𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹:(↑,↓)) and the polarization \n(linear: 𝜎𝜎=0 (black dots), and circ ular 𝜎𝜎±=±1) of the laser pulse as indicated in the legend \nand in the main text: (b) corresponds to 𝑥𝑥 =0.25 and (d) to x =0.24. MCD results are shown by \nsolid dots, whereas IFE results are presented by open symbols. Lines are guide to the eyes. \n \nFor 𝑥𝑥=0.25, as all result s presented up to here, the FiM is dominated by the RE:Gd \nat room temperature: 𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹<𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁 at 𝑇𝑇=300K , see FIG. 6(a) . In this case, the switching \nrequires less 𝑄𝑄 with circular polarization ( 𝜎𝜎±) with respect to the linearly polarized case \n19 \n (𝜎𝜎=0) for two different combinations of the circular laser polarization and the initial \nstate of the FiM: ( 𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↑) and (𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↓). This happens for both MCD (solid \nsymbols) and IFE scenarios (open symbols) as it is shown in FIG. 6(b) . Note that the \ninitial state in the TM is the opposite to the RE: 𝑚𝑚 𝑧𝑧𝑇𝑇𝐹𝐹↑ (up) corresponds to 𝑚𝑚 𝑧𝑧𝑅𝑅𝑁𝑁↓ (down) \nand vice, and the AOS is independent on the initial state for linear polarization (HI -AOS), \nwhereas under circular polarized laser pulses the switching is helicity -dependent (HD -\nAOS). For the rest of combinations, either (𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↑) or (𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↓), a higher 𝑄𝑄 is \nneed ed to achieve 100% of switching probability with respect to the linear polarized laser \npulse, and again for x =0.25 both MCD and IFE scenarios result in similar behavior of the \nswitching probability ( FIG. 6(b) ). \nRemarkably, the MCD and IFE r esults are qualitatively different w hen the \ncomposition is slightly modified to x =0.24, where the FiM becomes dominat ed by the \nTM:CoFe at room temperature: 𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹>𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁 at 𝑇𝑇=300K , see FIG. 6(c) . In this case \n(x=0.24) , the results in the IFE scenario are qualitatively similar to the ones already \nobtained for 𝑥𝑥=0.25: the HD -AOS occurs with small 𝑄𝑄 with respect to the linearly \npolarization case for (𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↑) and (𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↓) (open blue symbols in FIG. 6(d) ), and \nit requires high 𝑄𝑄 for the two other combinations ( (𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↑), (𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹↓), open red \nsymbols in FIG. 6(d) ). However, for this concentration ( x=0.24), the results in the MCD \nscenario (see solid symbols in FIG. 6(d) ) are opposite as for x =0.25, and also opposite to \nthe ones obtained in the IFE scenario. \nThese results can be understood as follows. In the MCD scenario, the HD -AOS \ndepends on the net initial magnetization at room temperature ( 𝑚𝑚𝑁𝑁=𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹+\n𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁, with 𝑚𝑚 𝑧𝑧𝑇𝑇𝐹𝐹=±1 and 𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁=∓1) and on the laser helicity ( 𝜎𝜎±=±1): if \ninitially 𝑚𝑚𝑁𝑁>0, a laser pulse with 𝜎𝜎+=+1 promotes the reversal, and this happens \neither for 𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹=−1 (𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁=+1) when x=0.25 because 𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁>𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹 at 𝑇𝑇=300K , or \nfor 𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹=+1 (𝑚𝑚𝑧𝑧𝑅𝑅𝑁𝑁=−1) if x=0.24 because now 𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹>𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁 at 𝑇𝑇=300K . On the \nother hand, the HD -AOS within the IFE scenario is essentially determined by the \ndominant sublattice magnetization just below the Curie threshold (𝑇𝑇≲𝑇𝑇𝐶𝐶), due to the \npersistence of the magneto -optical field 𝐵𝐵 �⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) when the laser puls e has already \nfinished. Note that for both concentrations 𝑀𝑀𝑠𝑠𝑇𝑇𝐹𝐹>𝑀𝑀𝑠𝑠𝑅𝑅𝑁𝑁 for 𝑇𝑇≲𝑇𝑇𝐶𝐶. Indeed, during the \ncooling down after the pulse, the magneto- optical field 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)∝𝜎𝜎±𝑢𝑢�⃗𝑧𝑧 promotes up \nor down magnetic state for the TM for 𝜎𝜎+ and 𝜎𝜎− respectively, and therefore, if 𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹 is 20 \n initially up (𝑚𝑚𝑧𝑧𝑇𝑇𝐹𝐹=+1), 𝐵𝐵�⃗𝐹𝐹𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) promotes the reversal for 𝜎𝜎− and vice. Our analysis \nsuggests a set of experiments which could help to elucidate the physical mechanism \nbehind these HD -AOS , just by evalu ating the switching probability as function of the \ninitial state and o f the laser pulse helicity for two different concentrations x , one resulting \nin a TM -dominated and other in a RE -dominated FiM alloy at room temperature. Another \nalternative could be to use a single FiM with a given composition, and working with a \ncryostat to fix different initial temperatures below and above the magnetization \ncompensation temperature. Similar results to ones obtained by changing the composition \nx for a fixed temperature of the thermal bath are also predicted by our simulations when \nit is the temperature of the thermal bath what is varied for a given composition (see Fig. \nS10 in Supplemental Material Note SN 9 [23] ): both IFE and MCD scenarios give similar \nresults below compensation and opposite above it . These theoretical predictions on the \nHD-AOS could be checked by experiments, which all together would allow us to shed \nlight on the real scope of these two mechanisms . \n \nIV. CONCLUSIONS \nAs summary, the extension of the two sublattice Landau- Lifshitz -Bloch equation \nwith the angular momentum non- equilibrium exchange is proven to be a powerful tool to \nstudy ultrafast AOS switching in ferrimagnetic alloys . The formalism here developed \nreproduces the atomistic spin dynamics results for small samples at the nanoscale , while \nit opens the possibility to numerically study realistic extended micro -size systems, with \ndimensions comparable to the experiment al one s. The deterministic single -shot switching \nand the demagnetization at high power regime are found to be in very good agreement with the experimental observations of Helicity -Independent AOS under linearly polarized \nlaser pulses . The phase diagram combined with the thermal analysis allow ed us to \ndetermine and predict the size of the inverted regions as depending the absorbed power \nand duration of the laser pulse . Moreover, w e have also explored and reproduced \nexperimental observations for the Helicity -Dependent AOS within the two physical \nmechanism s suggested in the literature: M agnetic Circular Dichroism and Inverse \nFaraday Effect. According to the Magnetic Circular Dichroism, the absorbed power by \nthe FiM depends on the laser helicity under circularly polarized pulses , and our model \nalso predicts the main features of the Helicity -Dependent AOS measurements. Indeed, 21 \n both the Helicity -Dependent AOS and the appearance of rings around the circularly \npolarized laser beam appear naturally in our simulations . Additionally, similar r esults of \nthe HD -AOS switching were also obtained in Inverse Faraday Effect scenario, where the \ncircular polarization has been suggested to generate a persistent magneto -optical field \npromoting the switching for proper combinations of initial magnetic stat e and laser pulse \nhelicity . By exploring FiM samples with different compositions resulting in TM -\ndominated or in a RE -dominated FiM alloy at room temperature, we have found a \ndifference between the predictions of the IFE and the MCD scenarios. These result s could \nbe tested by performing the corresponding experiments, and consequently helping \ntogether to elucidate the true basis of such HD-AOS processes . Therefore, our methods \nwill be useful to understand recent and future experiments on AOS , and a lso to the \ndevelop novel recording devices where the information can be manipulated by optical \nmeans in an ultra -fast fashion. \n \nACKNOWLEDGMENTS \nThis work was supported by projects MAT2017- 87072- C4-1-P funded by \nMinisterio de Educacion y Ciencia and PID2020117024GB -C41 funded by Ministerio de \nCiencia e Innovacion , both from the Spanish government , projects No. SA299P18 and \nSA114P20 from Consejer ia de Educaci on of Junta de Castilla y León, and project \nMagnEFi, Grant Agreement 860060, (H2020- MSCA -ITN-2019) funded by the European \nCommission. UA gratefully acknowledges support by the Deutsche \nForschungsgemeinschaft through SFB/TRR 227 \"Ultrafast Spin Dynamics\", Project A08. \n \n 22 \n REFERENCES \n[1] E. Beaurepaire, J. -C. Merle, A. Daunois, and J. -Y. Bigot, Ultrafast Spin \nDynamics in Ferromagnetic Nickel , Physical Review Letters 76, 4250 \n(1996). \n[2] M. S. el Hadri, P. Pirro, C. H. Lambert, S. Petit -Watelot, Y. Quessab, M. \nHehn, F. Montaigne, G. Malinowski, and S. Mangin, Two Types of All -\nOptical Magnetization Switching Mechanisms Using Femtosecond Laser \nPulses , Physical Review B 94, 064412 (2016). \n[3] R. Medapalli, D. Afanasiev, D. K. Kim, Y. Quessab, S. Manna, S. A. \nMontoya, A. Kirilyuk, T. Rasing, A. v. Kimel, and E. E. Fullerton, Multiscale \nDynamics of Heli city-Dependent All -Optical Magnetization Reversal in \nFerromagnetic Co/Pt Multilayers , Physical Review B 96, 224421 (2017). \n[4] M. Beens, M. L. M. Lalieu, A. J. M. Deenen, R. A. Duine, and B. Koopmans, \nComparing All -Optical Switching in Synthetic- Ferrimagne tic Multilayers \nand Alloys , Physical Review B 100 , 220409 (2019). \n[5] J. W. Liao, P. Vallobra, L. O’Brien, U. Atxitia, V. Raposo, D. Petit, T. Vemulkar, G. Malinowski, M. Hehn, E. Martínez, S. Mangin, and R. P. Cowburn, Controlling All- Optical Helicity -Dependent Switching in \nEngineered Rare -Earth Free Synthetic Ferrimagnets , Advanced Science 6, \n1901876 (2019). \n[6] M. L. M. Lalieu, M. J. G. Peeters, S. R. R. Haenen, R. Lavrijsen, and B. Koopmans, Deterministic All- Optical Switching of Synthetic Ferrimagnets \nUsing Single Femtosecond Laser Pulses , Physical Review B 96, 220411 \n(2017). \n[7] D. Steil, S. Alebrand, A. Hassdenteufel, M. Cinchetti, and M. Aeschlimann, \nAll-Optical Magnetization Recording by Tailoring Optical Excitation \nParameters , Physical Review B 84, 224408 (2011). \n[8] A. R. Khorsand, M. Savoini, A. Kirilyuk, A. v. Kimel, A. Tsukamoto, A. Itoh, \nand T. Rasing, Role of Magnetic Circular Dichroism in All -Optical Magnetic \nRecording, Physical Review Letters 108 , 127205 (2012). \n[9] K. Vahaplar, A. M. Kalash nikova, A. v. Kimel, S. Gerlach, D. Hinzke, U. \nNowak, R. Chantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing, All -\nOptical Magnetization Reversal by Circularly Polarized Laser Pulses: \nExperiment and Multiscale Modeling, Physical Review B 85, 104402 \n(2012). \n[10] M. O. A. Ellis, E. E. Fullerton, and R. W. Chantrell, All -Optical Switching in \nGranular Ferromagnets Caused by Magnetic Circular Dichroism , Scientific \nReports 6, 30522 (2016). \n[11] V. Raposo, E. Martinez, A. Hernandez, and M. Zazo, Micromagne tic \nModeling of All- Optical Switching, IEEE Transactions on Magnetics 55 , \n1300406 (2019). 23 \n [12] A. Kirilyuk, A. v. Kimel, and T. Rasing, Ultrafast Optical Manipulation of \nMagnetic Order , Reviews of Modern Physics 82, 2731 (2010). \n[13] V. Raposo, R. Guedas , F. García- Sánchez, M. A. Hernández, M. Zazo, and \nE. Martínez, Micromagnetic Modeling of All Optical Switching of Ferromagnetic Thin Films: The Role of Inverse Faraday Effect and Magnetic Circular Dichroism , Applied Sciences (Switzerland) 10 , 1307 \n(2020). \n[14] A. v. Kimel, A. Kirilyuk, P. A. Usachev, R. v. Pisarev, A. M. Balbashov, and T. Rasing, Ultrafast Non- Thermal Control of Magnetization by \nInstantaneous Photomagnetic Pulses , Nature 435, 655 (2005). \n[15] Lambert, C -H., S. Mangin, B. S. D. Ch. S. Varap rasad, Y. K. Takahashi, M. \nHehn, M. Cinchetti, G. Malinowski, K. Hono, Y. Fainman, M. Aeschlimann, \nand E. E. Fullerton, Ultrafast Optical Control of Orbital and Spin Dynamics \nin a Solid -State Defect , Science 345 , 1337 (2014). \n[16] T. A. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell, U. Atxitia, O. \nChubykalo -Fesenko, S. el Moussaoui, L. le Guyader, E. Mengotti, L. J. \nHeyderman, F. Nolting, A. Tsukamoto, A. Itoh, D. Afanasiev, B. A. Ivanov, \nA. M. Kalashnikova, K. Vahaplar, J. Mentink, A. Kirilyuk, T. Ras ing, and A. \nv. Kimel, Ultrafast Heating as a Sufficient Stimulus for Magnetization \nReversal in a Ferrimagnet , Nature Communications 3 , 1666 (2012). \n[17] I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius, H. A. Dürr, T. A. \nOstler, J. Barker, R. F. L. Ev ans, R. W. Chantrell, A. Tsukamoto, A. Itoh, A. \nKirilyuk, T. Rasing, and A. v. Kimel, Transient Ferromagnetic- like State \nMediating Ultrafast Reversal of Antiferromagnetically Coupled Spins , \nNature 472, 205 (2011). \n[18] C. S. Davies, T. Janssen, J. H. Menti nk, A. Tsukamoto, A. v. Kimel, A. F. G. \nvan der Meer, A. Stupakiewicz, and A. Kirilyuk, Pathways for Single -Shot \nAll-Optical Switching of Magnetization in Ferrimagnets , Physical Review \nApplied 13, 024064 (2020). \n[19] R. Moreno, T. A. Ostler, R. W. Chantrell, and O. Chubykalo -Fesenko, \nConditions for Thermally Induced All -Optical Switching in Ferrimagnetic \nAlloys: Modeling of TbCo , Physical Review B 96 , 014409 (2017). \n[20] P. Nieves, U. Atxitia, R. W. Chantrell, and O. Chubykalo -Fesenko, The \nClassical Two -Sublattice Landau- Lifshitz -Bloch Equation for All \nTemperatures , Low Temperature Physics 41, 739 (2015). \n[21] C. Vogler, C. Abert, F. Bruckner, and D. Suess, Stochastic Ferrimagnetic \nLandau- Lifshitz -Bloch Equation for Finite Magnetic Structures , Physic al \nReview B 100 , 054401 (2019). \n[22] R. F. L. Evans, W. J. Fan, P. Chureemart, T. A. Ostler, M. O. A. Ellis, and R. W. Chantrell, Atomistic Spin Model Simulations of Magnetic \nNanomaterials , Journal of Physics Condensed Matter. 26, 103202 (2014) 24 \n [23] See Supplemental Material at [URL will be inserted by publisher] for \n(SN1) Atomistic Spin Dynamics (ASD) model; (SN2) Micromagnetic LLB \nmodels: conventional (LLB) and extended (eLLB) cases; (SN3) Comparison between atomistic, conventional -LLB and extended -LLB models; (SN4). \nPhase diagram in terms of the fluence and the pulse duration; (SN5) Helicity -Dependent AOS (HD -AOS) and Magnetic Circular Dichroism \n(MCD); (SN6) Helicity -Dependent AOS (HD -AOS) and Inverse Faraday \nEffect (IFE); (SN7). Inverse Faraday Effect: magneto -optical field or \ninduced magnetic moment; (SN8) Material inputs for two different compositions; and (SN9) Helicity -Dependent All Optical Switching: MCD & \nIFE for different compositions and initial temperatures . \n[24] S. I. Anisimov, B. L. Kapeliovi ch, T. L. Perel’man, and L. D. Landau, Electron \nEmission from Metal Surfaces Exposed to Ultrashort Laser Pulses, Sov. J. Exp. Theor. Phys. 39, 375 (1974). \n[25] J. Mendil, P. Nieves, O. Chubykalo -Fesenko, J. Walowski, T. Santos, S. \nPisana, and M. Münzenberg, Resolving the Role of Femtosecond Heated \nElectrons in Ultrafast Spin Dynamics , Scientific Reports 4, 3980 (2014). \n[26] U. Atxitia, O. Chubykalo -Fesenko, J. Walowski, A. Mann, and M. \nMünzenberg, Evidence for Thermal Mechanisms in Laser -Induced \nFemtosecond Spin Dynamics , Physical Review B - Condensed Matter and \nMaterials Physics 81, 174401 (2010). \n[27] S. Gerlach, L. Oroszlany, D. Hinzke, S. Sievering, S. Wienholdt, L. \nSzunyogh, and U. Nowak, Modeling Ultrafast All -Optical Switching in \nSynthetic Ferrimagnet s, Physical Review B 95 , 224435 (2017). \n[28] C. Banerjee, N. Teichert, K. E. Siewierska, Z. Gercsi, G. Y. P. Atcheson, P. \nStamenov, K. Rode, J. M. D. Coey, and J. Besbas, Single Pulse All- Optical \nToggle Switching of Magnetization without Gadolinium in the \nFerrimagnet Mn2RuxGa, Nature Communications 11, 4444 (2020). \n[29] A. Ceballos, A. Pattabi, A. El -Ghazaly, S. Ruta, C. P. Simon, R. F. L. Evans, T. \nOstler, R. W. Chantrell, E. Kennedy, M. Scott, J. Bokor, and F. Hellman, \nRole of Element -Specific Damping in Ultrafast, Helicity- Independent, All -\nOptical Switching Dynamics in Amorphous (Gd,Tb)Co Thin Films , Physical \nReview B 103 , 024438 (2021). \n[30] U. Atxitia, P. Nieves, and O. Chubykalo -Fesenko, Landau- Lifshitz -Bloch \nEquation for Ferrimagnetic Materials , Physi cal Review B 86, 104414 \n(2012). \n[31] J. H. Mentink, J. Hellsvik, D. v. Afanasiev, B. A. Ivanov, A. Kirilyuk, A. v. Kimel, O. Eriksson, M. I. Katsnelson, and T. Rasing, Ultrafast Spin \nDynamics in Multisublattice Magnets , Physical Review Letters 108, \n057202 (2012). \n[32] J. H. Mentink, Manipulating Magnetism by Ultrafast Control of the \nExchange Interaction, Journal of Physics Condensed Matter. 29, 453001 \n(2017). 25 \n [33] V. G. Bar’yakhtar, V. I. Butrim, and B. A. Ivanov, Exchange Relaxation as a \nMechanism of the U ltrafast Reorientation of Spins in a Two -Sublattice \nFerrimagnet , JETP Letters 98, 289 (2013). \n[34] A. Kamra, R. E. Troncoso, W. Belzig, and A. Brataas, Gilbert Damping \nPhenomenology for Two -Sublattice Magnets , Physical Review B 98, \n184402 (2018). \n[35] J. Wei, B. Zhang, M. Hehn, W. Zhang, G. Malinowski, Y. Xu, W. Zhao, and S. Mangin, All -Optical Helicity -Independent Switching State Diagram in \nGd - Fe - Co Alloys , Physical Review Applied 15 , 054065 (2021). \n[36] S. Wang, C. Wei, Y. Feng, Y. Cao, H. Wang, W . Cheng, C. Xie, A. \nTsukamoto, A. Kirilyuk, T. Rasing, A. v. Kimel, and X. Li, All -Optical \nHelicity- Dependent Magnetic Switching by First -Order Azimuthally \nPolarized Vortex Beams , Applied Physics Letters 113, 171108 (2018). \n[37] C. Banerjee, K. Rode, G. Atcheson, S. Lenne, P. Stamenov, J. M. D. Coey, and J. Besbas, Ultrafast Double Pulse All -Optical Reswitching of a \nFerrimagnet , Physical Review Letters 126 , 177202 (2021). \n[38] M. Battiato, G. Barbalinardo, and P. M. Oppeneer, Quantum Theory of the Inverse F araday Effect , Physical Review B 89, 014413 (2014). \n[39] M. Berritta, R. Mondal, K. Carva, and P. M. Oppeneer, Ab Initio Theory of \nCoherent Laser -Induced Magnetization in Metals , Physical Review Letters \n117, 137203 (2016). \n \n 1 \n Supplementa l Material: Realistic micromagnetic description of all -optical ultrafast \nswitching processes in ferrimagnetic alloys \n \nV. Raposo1,*, F. García- Sánchez1, U. Atxitia2, and E. Martínez1,+ \n \n1. Applied Physics Department, University of Salamanca. \n2. Dahlem Center for Complex Quantum Systems and Fachbereich Physik . \n*,+ Corresponding authors : victor@usal.es , edumartinez@usal.es \n \n \nCONTENT: \n \n Supplemental Note SN1. Atomistic Spin Dynamics (ASD) model \n Supplemental Note SN2. Micromagnetic LLB model s: conventional (LLB) and \nextended (eLLB) cases \n Supplemental Note SN3. Comparison between atomistic, conventional -LLB \nand extended- LLB models \n Supplemental Note SN4. Phase diagram in terms of the fluence and the pulse \nduration \n Supplemental Note SN5. Helicity -Dependent AO S (HD-AOS) and M agnetic \nCircular Dichroism (MCD) \n Supplemental Note SN 6. Helicity -Dependent AO S (HD-AOS) and Inverse \nFaraday Effect ( IFE) \n Supplemental Note SN7. I nverse Faraday Effect: magneto-optical field or \ninduced magnetic moment \n Supplemental note SN 8. Material inputs for two different compositions \n Supplemental note SN 9: Helicity -Dependent All Optical Switching: MCD & \nIFE for different compositions and initial temperatures \n \n \n 2 \n SN1. Atomistic S pin Dynamics (ASD) model \nWith the growing power of modern computers, numerical modeling has become an \nessential tool for scientific research, especially when handling systems as complex as \nmagnetic devices. For magnetic modeling, there are several methodological choices in \nterms o f dimensions and time scales. In this section we review the bases of the Atomistic \nSpin Dynamic (ASD) model used to study the magnetization dynamics in small nano -size \nferrimagnetic FiM samples submitted to ultra -short of laser pulses. The physical basis o f \nthe ASD model is the localization of unpaired electrons to atomic sites, leading to an \neffective local atomistic magnetic moment at site 𝑖𝑖 , (𝜇𝜇⃗𝑖𝑖), which is treated as a classical \nspin of fi xed length. The FiM alloys are composed of two sublattices: rare -earth (RE) and \ntransition -metal (TM). Typical examples of these FiM alloys are GdCo, GdFe , GdFe Co, \nTbFeCo , etc. We consider that the ordered TM alloy is represented by the fcc -type lattice , \nand t o simulate the amorphous character of t he TM -RE alloy, 𝑥𝑥𝑅𝑅𝑅𝑅⋅100% lattice sites are \nsubstituted randomly with RE magnetic moments. An example of a typical atomistic \narrangement is shown in Fig. S1 for a Gd x(FeCo) 1-x with 𝑥𝑥=𝑥𝑥𝑅𝑅𝑅𝑅=0.25 being the \nrelative composition of the sublattices (𝑥𝑥=𝑥𝑥𝑅𝑅𝑅𝑅, and 𝑥𝑥 𝑇𝑇𝑇𝑇=1−𝑥𝑥𝑅𝑅𝑅𝑅). Each atom, either \nRE (Gd) or TM (FeCo), has its local atomistic magnetic moment 𝜇𝜇⃗𝑖𝑖 at site 𝑖𝑖 . \n \nFIG. S1. Atomistic scheme showing a FiM Gd x(FeCo) 1-x alloy consisting on two sublattices of \nRE (Gd, red) and TM (CoFe , grey) magnetic moments, with 𝑥𝑥 representing the relative \ncomposition of each sublattice. The sample dimensions along the Cartesian coordinate’s \ndirections are (ℓ×ℓ×𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇) with 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇=5.6 nm. Atomistic simulations are restricted to samples \nat the na noscale (ℓ≈25 nm). The dimensions of an elementary computational cell ( Δ𝑥𝑥=Δ𝑦𝑦=\n3 nm) within the micromagnetic approach in shown at the bottom, along with the indication of \nthe lattice constant ( 𝑎𝑎=0.32 nm) for comparison. \n \n3 \n The basis of ASD model (see for instance [1,2] and references therein ) is a classical \nspin Hamilt onian based on the Heisenberg exchange formalism. The spin Hamiltonian ℋ \ntypically has the form: \n ℋ=ℋ𝑒𝑒𝑒𝑒𝑒𝑒+ℋ𝑎𝑎𝑎𝑎𝑖𝑖+ℋ𝑑𝑑𝑖𝑖𝑑𝑑+ℋ𝑎𝑎𝑑𝑑𝑑𝑑 (eS1) \nwith the terms on the RHS representing respectively exchange, anisotropy, dipolar and \nZeeman terms. The exchange energy for a system of interacting atomic moments is given by the expression \n ℋ𝑒𝑒𝑒𝑒𝑒𝑒=−�𝐽𝐽𝑖𝑖𝑖𝑖𝑆𝑆⃗𝑖𝑖·𝑆𝑆⃗𝑖𝑖\n𝑖𝑖≠𝑖𝑖 (eS2) \nwhere 𝐽𝐽𝑖𝑖𝑖𝑖 is the exchange interaction betwe en atomic sites 𝑖𝑖 and 𝑗𝑗, 𝑆𝑆⃗𝑖𝑖 is a unit vector \ndenoting the local spin moment direction ( 𝑆𝑆⃗𝑖𝑖=𝜇𝜇⃗𝑖𝑖/𝜇𝜇𝑠𝑠𝑖𝑖 with 𝜇𝜇𝑠𝑠𝑖𝑖=|𝜇𝜇⃗𝑠𝑠𝑖𝑖|) and 𝑆𝑆⃗𝑖𝑖 is the spin \nmoment direction of neighboring atoms. The sum in Eq. (eS2) is truncated to nearest neighbors only. As the FiM alloy is formed by two sublattices of magnetic moments, we \ncan split the exchange interaction ( 𝐽𝐽\n𝑖𝑖𝑖𝑖) between ferromagnetic intra- lattice (𝐽𝐽𝑖𝑖𝑖𝑖>0) and \nantiferromagnetic inter -lattice (𝐽𝐽𝑖𝑖𝑖𝑖<0) exchange interactions. In what follows, we adopt \nthe notation of 𝐽𝐽 𝑇𝑇𝑇𝑇−𝑇𝑇𝑇𝑇>0, 𝐽𝐽𝑅𝑅𝑅𝑅−𝑅𝑅𝑅𝑅>0 for intra -lattice interactions, and 𝐽𝐽𝑇𝑇𝑇𝑇−𝑅𝑅𝑅𝑅<0 for \nthe inter -lattice interactions. \nThe second term in Eq. (eS2) is the magnetic anisotropy. Here, a standard uniaxial \nanisotropy along the easy -axis (𝑧𝑧, out-of-plane direction ) is considered , \n ℋ𝑎𝑎𝑎𝑎𝑖𝑖=−𝑑𝑑𝑢𝑢��𝑆𝑆�⃗𝑖𝑖·𝑢𝑢�⃗𝐾𝐾�2\n𝑖𝑖 (eS3) \nwhere 𝑑𝑑𝑢𝑢 is the anisotropy energy per atom and 𝑢𝑢�⃗𝐾𝐾=𝑢𝑢�⃗𝑧𝑧 is the unit vector denoting the \npreferred direction . The last two terms in Eq. ( eS1) account for the dipolar (ℋ𝑑𝑑𝑖𝑖𝑑𝑑) and \nexternal applied field (ℋ𝑎𝑎𝑑𝑑𝑑𝑑) contributions . Since the demagnetizing field is usually \nrelatively small in FiMs samples , this term is generally ignored here. No external field is \napplied in the present work. \nThe equilibrium state of the FiM sample can be obtained by minimiz ing of the total \nenergy. Here, it consists o f the two antiferromagnetic coupled sublattices aligned along \nthe easy -axis in anti -parallelly . Its dynamics response is governed by the Langevin -\nLaudau -Lifshitz -Gilbert equation for each atomistic moment ( 𝑆𝑆⃗𝑖𝑖) [1,2] : \n𝜕𝜕𝑆𝑆⃗𝑖𝑖\n𝜕𝜕𝑡𝑡=−𝛾𝛾0\n(1+𝜆𝜆2)�𝑆𝑆⃗𝑖𝑖×�𝐻𝐻��⃗𝑖𝑖+𝐻𝐻��⃗𝑡𝑡ℎ,𝑖𝑖 �+𝜆𝜆𝑆𝑆⃗𝑖𝑖×�𝑆𝑆⃗𝑖𝑖×�𝐻𝐻��⃗𝑖𝑖+𝐻𝐻��⃗𝑡𝑡ℎ,𝑖𝑖 ��� (eS4) \nwhere 𝜆𝜆=0.02 is the Gilbert damping parameter and 𝛾𝛾0=2.21×105 m/(A⋅s) the \ngyromagnetic ratio . 𝐻𝐻��⃗𝑖𝑖 is the local effective magnetic field obtained from the spin \nHamiltonian as \n𝜇𝜇0𝐻𝐻��⃗𝑖𝑖=−1\n𝜇𝜇𝑠𝑠𝑖𝑖𝜕𝜕ℋ\n𝜕𝜕𝑆𝑆⃗𝑖𝑖 (eS5) 4 \n and 𝐻𝐻��⃗𝑡𝑡ℎ,𝑖𝑖 is the stochastic thermal field is given by: \n𝜇𝜇0𝐻𝐻��⃗𝑡𝑡ℎ,𝑖𝑖 =Γ⃗𝑖𝑖(𝑡𝑡)�2𝜆𝜆𝑘𝑘𝐵𝐵𝑇𝑇\n𝛾𝛾0𝜇𝜇𝑠𝑠𝑖𝑖Δ𝑡𝑡 (eS6) \nwith 𝑘𝑘𝐵𝐵 is the Boltzmann constant, 𝑇𝑇 the temperature, Δ 𝑡𝑡 the integration time step . Γ⃗𝑖𝑖(𝑡𝑡) \nis a local vector wh ose components a re Gaussian -distributed white -noise random \nnumbers with zero mean value . \nUnder a laser pulse, the FiM sample absorbs energy and its temperature changes in time. \nThe temperature 𝑇𝑇 that enters in Eq. (eS6) is the temperature of the electronic subsystem \n(𝑇𝑇≡𝑇𝑇𝑒𝑒), which is coupled to the lattice subsystem temperature ( 𝑇𝑇𝑙𝑙) and to the laser pulse \nas given b y the Two Temperature Model ( TTM ). Details of the power absorbed by sample \nas due to the laser pulse and the Eqs. (2)-(3) of the TTM were already discussed in the \nmain text. Conventional TTM values were adopted [3 –5]: 𝐶𝐶𝑒𝑒=1.8×105 J/(m3K) at \n𝑇𝑇𝑅𝑅= 300 K, 𝐶𝐶𝑙𝑙=3.8×106 J/(m3K), 𝑘𝑘𝑒𝑒=91 W/(m K), 𝑔𝑔𝑒𝑒𝑙𝑙=7×1017 W/(m3K) \nand 𝜏𝜏𝐷𝐷=10 ps. \nEq. (eS4) [or Eq. (1) in the main text] is numerically solved coupled to the T TM \nEqs. ((2) -(3) in the main text) with a home -made solver using a 4th-order Runge -Kutta \nscheme with Δ𝑡𝑡=0.1 fs. Typical Gd 0.25(CoFe )0.75 parameters were considered [2]: intra-\nlattice exchange energies: intra-lattice exchange energies 𝐽𝐽 𝑇𝑇𝑇𝑇−𝑇𝑇𝑇𝑇=3.58×10−21 J, \n𝐽𝐽𝑅𝑅𝑅𝑅−𝑅𝑅𝑅𝑅=1.44×10−21 J, inter -lattice energy 𝐽𝐽𝑇𝑇𝑇𝑇−𝑅𝑅𝑅𝑅=−1.25×10−21 J, magnitude of \nthe local magnetic moments: 𝜇𝜇𝑇𝑇𝑇𝑇=1.92 𝜇𝜇𝐵𝐵, 𝜇𝜇𝑅𝑅𝑅𝑅=7.63 𝜇𝜇𝐵𝐵, and anisotropy energy of \n𝑑𝑑𝑢𝑢=8.07×10−24 J. The FiM sample dimensions for atomistic simulations are \nℓ×ℓ×𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 (see FIG. S1) with ℓ ≃25 nm and 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 =5.6 nm, and magnetic moment \nof the two sublattices are randomly generated with 25% and 75% of RE:Gd and \nTM:(FeCo) respectively (FIG. S1). The lattice constant is 𝑎𝑎 =0.35 nm . TTM parameter \nThe power absorbed by the laser pulse is assumed to be uniform over the nano- size FiM \nsample considered in the main text for atomistic simulations (ASD) . \n \nSN2. Micromagnetic LLB models: conventional (LLB) and extended (eLLB) cases \nAs it was already stated, due to memory and computation limitations , atomistic \nsimulations are restricted to small samples at the nano -scale (ℓ≲100 nm ). However, \nexperimental studies on All Optical Switching are typical carried out in extended samples \nwith several micro meters in length ( ℓ≃100 μm ). Also, the size of conventional laser \nspot (FWHM) are usually at the micro -scale (𝑑𝑑0~1−10 μm). It is worthy to provide an \nestimation of the maximum size that c ould be numerically manage d in atomistic \nsimulations (ASD) . For instance, in the present work we used one of the most power ful \nGraphics Processing Units (GPUs) , model RTX3090 (24GB) , and performing atomistic \nsimulations of FiM samples of ℓ×ℓ×𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 with ℓ≃150 nm and 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇=5.6 nm \n(𝑎𝑎~0.35 nm ) will be already out of the computational power of such GPU. Therefore, in \norder to explain experimental observations another coarse- grained formalism is needed. \nHere, we adopt the mesoscopic description ( micromagnetic model) which assumes that \nthe magnetization of each sublattice is a continuous vector field: 𝑚𝑚��⃗𝑖𝑖=𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡), where 5 \n here the sub- index 𝑖𝑖 refers to the local magnetic magnetization of each sublattice, \n𝑖𝑖:RE,TM. The FiM sample, with dimension ℓ×ℓ×𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇, is discretized using a 2D finite \ndifferences scheme, using computational cells of volume Δ 𝑉𝑉=Δ𝑥𝑥2 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇. The continuous \ndescription ( Δ𝑥𝑥≫𝑎𝑎) is justified by the short- range of the exchange interaction, which \npromote s the ferromagnetic and the antiferromagnetic orders for intra - and inter -lattices \ninteractions respectively. At the same time, in ord er to resolve magnetic patterns such as \ndomain walls, the cell size Δ𝑥𝑥 must be smaller than the characteristic length scale along \nwhich the magnetization varies significantly (for instance, the so -called domain wall \nwidth , 𝛿𝛿≫Δ𝑥𝑥). Therefore, within the m icromagnetic approach, at each cell location ( 𝑟𝑟⃗) \nthere are two magnetic spin s representing the magnetization of the two sublattices of the \nFiM, 𝑚𝑚��⃗𝑖𝑖=𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)=𝑀𝑀��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)/𝑀𝑀𝑠𝑠𝑖𝑖(𝑇𝑇) where 𝑀𝑀��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡) is the local magnetization of \nsublattice 𝑖𝑖 in units of A/m, and 𝑀𝑀 𝑠𝑠𝑖𝑖(𝑇𝑇) is the corresponding spontaneous magnetization \nat temperature 𝑇𝑇. FIG. S2 shows these ideas, along with different spatial scales of the \nASD and micromagnetic model. \n \n \nFIG. S2 . (a) Arrange of atoms for atomistic simulations , limited to nanoscale samples, ℓ≈\n25 nm. (b) Micromagnetic discretization scheme for extended samples at the microscale, ℓ ≈\n20 μm. The FiM sample is discretized using a 2D finite differences scheme using computational \ncells of volume 𝑉𝑉 =Δ𝑥𝑥2 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇. Each cell contains two micro -magnetic moments, one for each \nsublattice , and its size would include thousands of atomistic magnetic moments (𝑆𝑆⃗𝑖𝑖). The \nmagnetization of each sublattice i s a continuous vector function (𝑚𝑚��⃗𝑖𝑖=𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)) over the FiM \nsample. \n \nIn order to overcome the ASD limitations here we adopt an extended continuous \nmicromagnetic model that describes the temporal evolution of the reduced local \nmagnetization 𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡) of each sublattice 𝑖𝑖 :RE,TM based on the stochastic Landau -\nLifshitz -Bloch (LLB) equation [1,6] \n \n6 \n 𝜕𝜕𝑚𝑚��⃗𝑖𝑖\n𝜕𝜕𝑡𝑡=−𝛾𝛾0𝑖𝑖′�𝑚𝑚��⃗𝑖𝑖×𝐻𝐻��⃗𝑖𝑖�+ \n−𝛾𝛾0𝑖𝑖′𝛼𝛼𝑖𝑖⊥\n𝑚𝑚𝑖𝑖2𝑚𝑚��⃗𝑖𝑖×�𝑚𝑚��⃗𝑖𝑖×�𝐻𝐻��⃗𝑖𝑖+𝜉𝜉⃗𝑖𝑖⊥��− \n+𝛾𝛾0𝑖𝑖′𝛼𝛼𝑖𝑖∥\n𝑚𝑚𝑖𝑖2(𝑚𝑚��⃗𝑖𝑖·𝐻𝐻��⃗𝑖𝑖)𝑚𝑚��⃗𝑖𝑖+𝜉𝜉⃗𝑖𝑖∥ (eS7) \nwhere 𝛾𝛾0𝑖𝑖′=𝛾𝛾0𝑖𝑖/(1 +𝜆𝜆𝑖𝑖2) is the reduced gyromagnetic ratio, which is defined via the \ncoupling parameter 𝜆𝜆𝑖𝑖 of sublattice 𝑖𝑖 to the heat bath . 𝛼𝛼𝑖𝑖∥ and 𝛼𝛼𝑖𝑖⊥ are the longitudinal and \nperpendicular damping parameters . 𝐻𝐻��⃗𝑖𝑖=𝐻𝐻��⃗𝑖𝑖(𝑟𝑟 ⃗,𝑡𝑡) is the local effective field at location 𝑟𝑟 ⃗ \nacting on sublattice magnetic moment 𝑖𝑖, and 𝜉𝜉⃗𝑖𝑖∥ and 𝜉𝜉⃗𝑖𝑖⊥ are the longitudinal and \nperpendicular stochastic thermal fields. \nFor 𝑇𝑇<𝑇𝑇𝐶𝐶, the damping constants for sublattice 𝑖𝑖 are \n𝛼𝛼𝑖𝑖∥= 2𝜆𝜆𝑖𝑖𝑘𝑘𝐵𝐵𝑇𝑇𝑚𝑚𝑒𝑒,𝑖𝑖\n𝐽𝐽0,𝑖𝑖𝑖𝑖𝑚𝑚𝑒𝑒,𝑖𝑖+�𝐽𝐽0,𝑖𝑖𝑖𝑖�𝑚𝑚𝑒𝑒,𝑖𝑖 (eS8) \n𝛼𝛼𝑖𝑖⊥=𝜆𝜆𝑖𝑖�1−𝑘𝑘𝐵𝐵𝑇𝑇𝑚𝑚𝑒𝑒,𝑖𝑖\n𝐽𝐽0,𝑖𝑖𝑖𝑖𝑚𝑚𝑒𝑒,𝑖𝑖+�𝐽𝐽0,𝑖𝑖𝑖𝑖�𝑚𝑚𝑒𝑒,𝑖𝑖� (eS9) \nwhere 𝑚𝑚𝑒𝑒,𝑖𝑖 and 𝑚𝑚𝑒𝑒,𝑖𝑖 are the equilibrium magnetization in sublattices 𝑇𝑇 𝑀𝑀 a nd 𝑅𝑅𝑅𝑅 \nrespectively . These values are obtained from the mean field approximation (see \nRef. [1,6]) . The exchange energies became 𝐽𝐽0,𝑖𝑖𝑖𝑖=𝑧𝑧𝑥𝑥𝑖𝑖𝐽𝐽𝑖𝑖−𝑖𝑖, 𝐽𝐽0,𝑖𝑖𝑖𝑖=𝑧𝑧𝑥𝑥𝑖𝑖𝐽𝐽𝑖𝑖−𝑖𝑖, with 𝑧𝑧 the \nnumber of nearest neighbors and 𝐽𝐽𝑖𝑖−𝑖𝑖 and 𝐽𝐽𝑖𝑖−𝑖𝑖 the atomistic exchange energy between \nspecies. 𝑥𝑥𝑖𝑖 is the concentration of sublattice 𝑖𝑖 . Above the Curie temperature, 𝑇𝑇>𝑇𝑇𝐶𝐶, the \ndamping parameters are \n𝛼𝛼𝑖𝑖∥=𝛼𝛼𝑖𝑖⊥=2𝜆𝜆𝑖𝑖𝑇𝑇\n3𝑇𝑇𝐶𝐶 (eS10 ) \nThe effective field consists on several contributions: \n𝐻𝐻��⃗𝑖𝑖=𝐻𝐻��⃗𝑎𝑎𝑎𝑎𝑖𝑖,𝑖𝑖+𝐻𝐻��⃗𝑒𝑒𝑒𝑒,𝑖𝑖+𝐻𝐻��⃗𝑖𝑖∥ (eS11 ) \nwhere 𝐻𝐻��⃗𝑎𝑎𝑎𝑎𝑖𝑖,𝑖𝑖 and 𝐻𝐻��⃗𝑒𝑒𝑒𝑒,𝑖𝑖 are the anisotropy and exchange fields given by \n𝐻𝐻��⃗𝑎𝑎𝑎𝑎𝑖𝑖,𝑖𝑖=2𝐾𝐾𝑖𝑖\n𝜇𝜇0𝜇𝜇𝑖𝑖𝑚𝑚��⃗𝑖𝑖·𝑢𝑢�⃗𝐾𝐾 (eS12 ) \n𝐻𝐻��⃗𝑒𝑒𝑒𝑒,𝑖𝑖=2𝐴𝐴𝑒𝑒𝑒𝑒,𝑖𝑖\n𝜇𝜇0𝑀𝑀𝑠𝑠𝑖𝑖(∇2𝑚𝑚��⃗𝑖𝑖)−𝐽𝐽0,𝑖𝑖𝑖𝑖\n𝜇𝜇0𝜇𝜇𝑖𝑖𝑚𝑚𝑖𝑖2�𝑚𝑚��⃗𝑖𝑖× (𝑚𝑚��⃗𝑖𝑖×𝑚𝑚��⃗𝑖𝑖)� (eS13 ) \nwhere the first term in 𝐻𝐻��⃗𝑒𝑒𝑒𝑒,𝑖𝑖 is the continuous exchange between neighbors’ cells, and the \nsecond one accounts for the inter -lattice contribution. Finally, the term 𝐻𝐻��⃗𝑖𝑖∥ is a f ield \ncomputed differently below 𝑇𝑇 𝐶𝐶, \n𝐻𝐻��⃗𝑖𝑖∥(𝑇𝑇<𝑇𝑇𝐶𝐶) =−1\n𝜇𝜇0�1\n𝜒𝜒 �𝑖𝑖∥+�𝐽𝐽0,𝑖𝑖𝑖𝑖�\n𝜇𝜇𝑖𝑖𝜒𝜒 �𝑖𝑖∥\n𝜒𝜒 �𝑖𝑖∥�𝛿𝛿𝑚𝑚𝑖𝑖\n𝑚𝑚𝑒𝑒,𝑖𝑖 𝑚𝑚��⃗𝑖𝑖+1\n𝜇𝜇0�𝐽𝐽0,𝑖𝑖𝑖𝑖�\n𝜇𝜇𝑖𝑖𝛿𝛿𝜏𝜏𝐵𝐵,𝑖𝑖\n𝑚𝑚𝑒𝑒,𝑖𝑖𝑚𝑚��⃗𝑖𝑖 (eS14 ) 7 \n and above 𝑇𝑇𝐶𝐶 \n𝐻𝐻��⃗𝑖𝑖∥(𝑇𝑇 ≥ 𝑇𝑇𝐶𝐶) =−1\n𝜇𝜇0�1\n𝜒𝜒 �𝑖𝑖∥+�𝐽𝐽0,𝑖𝑖𝑖𝑖�\n𝜇𝜇𝑖𝑖𝜒𝜒�𝑖𝑖∥\n𝜒𝜒 �𝑖𝑖∥�𝑚𝑚��⃗𝑖𝑖+1\n𝜇𝜇0�𝐽𝐽0,𝑖𝑖𝑖𝑖�\n𝜇𝜇𝑖𝑖𝜏𝜏𝐵𝐵,𝑖𝑖\n𝑚𝑚𝑖𝑖𝑚𝑚��⃗𝑖𝑖 (eS15 ) \nwith 𝜏𝜏𝐵𝐵,𝑖𝑖=�𝑚𝑚��⃗𝑖𝑖·𝑚𝑚��⃗𝑖𝑖�/𝑚𝑚𝑖𝑖, 𝛿𝛿𝑚𝑚𝑖𝑖=𝑚𝑚𝑖𝑖−𝑚𝑚𝑒𝑒,𝑖𝑖 and 𝛿𝛿𝜏𝜏𝐵𝐵,𝑖𝑖=𝜏𝜏𝐵𝐵,𝑖𝑖−𝜏𝜏𝑒𝑒,𝐵𝐵,𝑖𝑖. The longitudinal \nsusceptibilities 𝜒𝜒 �𝑖𝑖∥ are calculated from mean field approach [1] . Below the Curie \ntemperature ( 𝑇𝑇<𝑇𝑇𝐶𝐶): \n𝜒𝜒 �𝑖𝑖∥(𝑇𝑇<𝑇𝑇𝐶𝐶)\n=𝜇𝜇𝑖𝑖ℒ𝑖𝑖′(𝜁𝜁𝑖𝑖)�𝐽𝐽0,𝑖𝑖𝑖𝑖�ℒ𝑖𝑖′�𝜁𝜁𝑖𝑖�+𝜇𝜇𝑖𝑖ℒ𝑖𝑖′(𝜁𝜁𝑖𝑖)�𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖ℒ𝑖𝑖′�𝜁𝜁𝑖𝑖��\n�𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖ℒ𝑖𝑖′(𝜁𝜁𝑖𝑖)��𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖ℒ𝑖𝑖′�𝜁𝜁𝑖𝑖��−�𝐽𝐽0,𝑖𝑖𝑖𝑖�ℒ𝑖𝑖′(𝜁𝜁𝑖𝑖)�𝐽𝐽0,𝑖𝑖𝑖𝑖�ℒ𝑖𝑖′�𝜁𝜁𝑖𝑖� (eS16 ) \nwhere ℒ𝑖𝑖 is the Langevin function with argument 𝜁𝜁𝑖𝑖=𝐽𝐽0,𝑖𝑖𝑖𝑖𝑚𝑚𝑖𝑖+�𝐽𝐽0,𝑖𝑖𝑖𝑖�𝑚𝑚𝑖𝑖\n𝑘𝑘𝐵𝐵𝑇𝑇 and ℒ𝑖𝑖′ is the \nderivative respect 𝜁𝜁𝑖𝑖. Above the Curie temperature ( 𝑇𝑇>𝑇𝑇𝐶𝐶): [7] \n𝜒𝜒 �𝑖𝑖∥(𝑇𝑇 ≥ 𝑇𝑇𝐶𝐶) =𝜇𝜇𝑖𝑖�𝐽𝐽0,𝑖𝑖𝑖𝑖�+𝜇𝜇𝑖𝑖�3𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖�\n��3𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖��3𝑘𝑘𝐵𝐵𝑇𝑇−𝐽𝐽0,𝑖𝑖𝑖𝑖�−�𝐽𝐽0,𝑖𝑖𝑖𝑖��𝐽𝐽0,𝑖𝑖𝑖𝑖�� (eS17) \n \nStochastic fluctuations due to temperature are introduced through thermal fields 𝜉𝜉 ⃗𝑖𝑖∥ \nand 𝜉𝜉⃗𝑖𝑖⊥. These fields are time and space uncorrelated, and they are generated from white \nnoise random numbers with zero mean field and variance given by [6]: \n〈𝜉𝜉𝑖𝑖,𝛼𝛼𝜂𝜂(𝑟𝑟⃗,𝑡𝑡)𝜉𝜉𝑖𝑖,𝛽𝛽𝜂𝜂(𝑟𝑟⃗′,𝑡𝑡′)〉=2𝐷𝐷𝑖𝑖𝜂𝜂𝛿𝛿𝛼𝛼𝛽𝛽𝛿𝛿(𝑟𝑟⃗−𝑟𝑟⃗′)𝛿𝛿(𝑡𝑡−𝑡𝑡′) (eS18) \nwhere 𝛼𝛼,𝛽𝛽 are the cartesian components of the stochastic thermal fields , 𝑖𝑖 denotes \nsublattice 𝑇𝑇\n𝑀𝑀 o\nr 𝑅𝑅𝑅𝑅 and 𝜂𝜂:∥,⊥ represents parallel or perpendicular field components. \nThe diffusions constants are given by: \n𝐷𝐷𝑖𝑖⊥=�𝛼𝛼𝑖𝑖⊥−𝛼𝛼𝑖𝑖∥�𝑎𝑎3𝑘𝑘𝐵𝐵𝑇𝑇\n(𝛼𝛼𝑖𝑖⊥)2𝛾𝛾0𝑖𝑖′𝜇𝜇0𝑛𝑛𝑎𝑎𝑡𝑡𝑥𝑥𝑖𝑖𝜇𝜇𝑖𝑖Δ𝑉𝑉 (eS19) \n𝐷𝐷𝑖𝑖∥=𝛼𝛼𝑖𝑖∥𝛾𝛾0𝑖𝑖′𝑎𝑎3𝑘𝑘𝐵𝐵𝑇𝑇\n𝑛𝑛𝑎𝑎𝑡𝑡𝑥𝑥𝑖𝑖𝜇𝜇0𝜇𝜇𝑖𝑖Δ𝑉𝑉 (eS20) \nwith Δ𝑉𝑉=Δ𝑥𝑥2𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 the discretization volume, 𝑎𝑎 is the lattice constant and 𝑛𝑛𝑎𝑎𝑡𝑡 is the \nnumber of atoms per unit cell. \n \nA\ns it will be shown in Supplemental Note N 3(b), we have confirmed that the \nconventional LLB Eq. (eS7) (or Eq. (4) in the main text) [1,6,8] is not able to fully \nreproduce the atomistic model predictions of some all optical switching processes in FiM \nsystems. Indeed, atomistic simulations show that the sudden change in the temperature \ndue to the laser pulse results in transient non-equilibrium states where a transfer of angular \nmomentum between sublattices takes place. This transference is not fully reproduced by \nthe conventional LLB Eq. (eS7). In order to overcome this limitation, in the present work 8 \n we exten d the LLB E q. (eS7) by add ing an additional term to the RHS that account s of \nthe angular momentum transfer between lattices in non -equilibrium states . This non-\nequilibrium torque is given \n𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅=𝛾𝛾0𝑖𝑖′𝜆𝜆𝑒𝑒𝑒𝑒𝛼𝛼𝑖𝑖∥𝑥𝑥𝑖𝑖𝜇𝜇𝑖𝑖𝑚𝑚𝑖𝑖+𝑥𝑥𝑖𝑖𝜇𝜇𝑖𝑖𝑚𝑚𝑖𝑖\n𝜇𝜇𝑖𝑖𝑚𝑚𝑖𝑖𝜇𝜇𝑖𝑖𝑚𝑚𝑖𝑖�𝜇𝜇𝑖𝑖𝐻𝐻��⃗𝑖𝑖∥−𝜇𝜇𝑖𝑖𝐻𝐻��⃗𝑖𝑖∥� (eS21) \nwhere and 𝐻𝐻��⃗𝑖𝑖∥ and 𝐻𝐻��⃗𝑖𝑖∥ are the longitudinal effective fields for each lattice 𝑖𝑖 :RE,TM [6], \n𝑥𝑥𝑖𝑖≡𝑥𝑥 and 𝑥𝑥𝑖𝑖=1−𝑥𝑥𝑖𝑖≡1−𝑥𝑥 are the concentrations of each specimen, and 𝜆𝜆𝑒𝑒𝑒𝑒 is a \nparameter representing the exchange relaxation rate [9]. By including Eq. (eS21) in the \nRHS of Eq. (eS7), and numerically solving it coupled to TTM Eqs. (2) -(3), we can provide \na realistic description of the magnetization dynamics in FiM systems under ultra -short \nlaser pulses. In what follows, and in order to distinguish it from the conventional LLB \nmodel (LLB, 𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅=0), we refer to this formalism as the extende d micromagnetic LLB \nmodel (eLLB, 𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅≠0), Note that the additional torque has not been previously included \nin the LLB model, and therefore, here we name it here as the extended LLB model \n(eLLB) . A comparison between atomistic (ASD) predictions and the results without \n(conventional LLB) and with (extended LLB, eLLB) the additional t orque 𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅 is shown \nin FIG. S4, discussed in the next section. \nIn the present work , FiM samples of ℓ×ℓ×𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 are discretized using a 2D finite \ndifferences scheme, using computational cells of volume Δ 𝑉𝑉=Δ𝑥𝑥2 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇. Typical \nmaterial parameters for Gd 0.25(FeCo) 0.75 were adopted: 𝐽𝐽0,𝑇𝑇𝑇𝑇−𝑇𝑇𝑇𝑇=2.59×10−21 J, \n 𝐽𝐽0,𝑅𝑅𝑅𝑅−𝑅𝑅𝑅𝑅=1.35×10−21 J, 𝐽𝐽0,𝑇𝑇𝑇𝑇−𝑅𝑅𝑅𝑅=−1.13×10−21 J. 𝜆𝜆𝑇𝑇𝑇𝑇=𝜆𝜆𝑅𝑅𝑅𝑅=0.02. \nMicromagnetic parameters needed to solve the LLB Eq. can be obtained from atomistic \ninputs as follows [2]: the spontaneous magnetization at zero temperature of each \nsublattice is 𝑀𝑀 𝑠𝑠𝑖𝑖(0)=𝜇𝜇𝑖𝑖𝑥𝑥𝑖𝑖/(𝑝𝑝𝑓𝑓⋅𝑎𝑎3), where 𝑝𝑝𝑓𝑓 is a packing factor which depends on the \ncrystalline structure ( 𝑝𝑝𝑓𝑓=0.74 for fcc, [2]): 𝑀𝑀𝑆𝑆,𝑇𝑇𝑇𝑇(0)=0.41 MA /m, 𝑀𝑀𝑆𝑆,𝑅𝑅𝑅𝑅(0)=\n0.55 MA /m. The continuous exchange stiffness constant [10] is 𝐴𝐴𝑒𝑒𝑒𝑒,𝑖𝑖=𝑛𝑛𝑛𝑛𝑥𝑥𝑖𝑖2𝐽𝐽𝑖𝑖/(2𝑎𝑎) \nwith 𝑛𝑛=2 being the number of atoms pe r unit cell for fcc, 𝑛𝑛=0.79 the spin wave mean \nfield correction value also for FCC, and 𝑥𝑥𝑖𝑖 the concentration of sublattice 𝑖𝑖 (𝑥𝑥𝑖𝑖=1−𝑥𝑥𝑖𝑖): \n𝐴𝐴𝑒𝑒𝑒𝑒,𝑇𝑇𝑇𝑇=3.2 pJ/m, 𝐴𝐴𝑒𝑒𝑒𝑒,𝑅𝑅𝑅𝑅=0.19 pJ /m. The perpendicular anisotropy parameter is \n𝐾𝐾𝑖𝑖=𝑑𝑑𝑢𝑢𝑥𝑥𝑖𝑖/(𝑝𝑝𝑓𝑓⋅𝑎𝑎3), see [2] , so 𝐾𝐾𝑢𝑢,𝑇𝑇𝑇𝑇=1.87 MJ /m3, 𝐾𝐾𝑢𝑢,𝑅𝑅𝑅𝑅=0.62 MJ /m3. For the \npresented results we take 𝜆𝜆𝑒𝑒𝑒𝑒=0.013 for the exchange relaxation rate. Within the \nmicromagnetic model, the laser -induced magnetization dynamics is evaluated by \nnumerically solving Eq. (eS7) coupled to the TTM Eqs. (2) -(3) using the Heun algorithm \nwith Δ𝑡𝑡=1 fs with cell sizes of Δ𝑥𝑥=3 nm. This was done by implementing the models \nin a home -made CUDA -based code, which was run a RTX3090 GPU . It was also checked \nin several tests that decreasing the cell size to Δ 𝑥𝑥=1 nm does not change the results. \n \nSN3. Comparison between atomistic, conventional LLB and extended LLB models \n(a) In order to validate the extended LLB model (eLLB, 𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅≠0), we have firstly \ncompared the predictions of the conventional LLB model (LLB , 𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅=0) with the \natomistic results (ASD) by computing for the equilibrium magnetization of each 9 \n sublattice ( 𝑖𝑖:RE,TM ) as a function of the temperature of the thermal bath (w hich coincides \nwith the electronic temperature). This study was carried out in a small FiM sample with \nℓ≈25 nm, and the results are shown in F IG. S3. Both ASD and eLLB models lead to \nsimilar results . \n \nFIG. S3. Temperature dependence of the reduced equilibrium magnetization (𝑚𝑚) of the two \nsublattices . Solid (dashed ) lines corresponds to ASD ( eLLB) results. Blue and red color \ncorrespond to the RE and TM sublattices respectively. The inset shows the corresponding \ndependence of the spontaneous magnetiz ation value of each sublattice ( 𝑀𝑀𝑠𝑠𝑖𝑖(𝑇𝑇) vs 𝑇𝑇). \n \n(b) In the main text and also in S N2, we claimed that the conventional LLB mode l \n(LLB , 𝜏𝜏⃗𝑁𝑁𝑅𝑅,𝑖𝑖=0) is not able to fully reproduce the atomistic results (ASD) for some all -\noptical switching processes in FiM samples. On the other hand, when such term is \nconsidered ( 𝜏𝜏⃗𝑁𝑁𝑅𝑅,𝑖𝑖≠0), the extended LLB model (eLLB) naturally reproduces the \natomistic temporal variation of the magnetization in small FiM samples under ultra -short \nlaser pulses. This is shown in F IG. S4 for the same study as in Fig. 1(b) of the main text. \nContrary to the ASD and the eLLB models, the conventional LLB (𝜏𝜏⃗𝑁𝑁𝑅𝑅,𝑖𝑖=0) does not \npredicts switching for 𝑄𝑄=12×1021 W/m3 pulses ( Fig. S4(c) -(d)). Moreover, as shown \nin FIG. S4(a) -(b), t he extended LLB model (eLLB) , which includes 𝜏𝜏⃗𝑁𝑁𝑅𝑅,𝑖𝑖≠0, repro duces \nsimilar results even in the absence of therma l fluctuations ( 𝜉𝜉⃗𝑖𝑖∥=𝜉𝜉⃗𝑖𝑖⊥=0). \n10 \n \nFIG. S4. Temporal evolution of the out -of-plane component (𝑚𝑚𝑧𝑧 vs 𝑡𝑡) calculated by three different \nmodels : ASD, conventional LLB and extended eLLB models for the case studied in Fig. 1(b) of \nthe main text. Solid lines in all graphs correspond to the ASD results. Red and blue c urves \ncorrespond to the TM and RE respectively . In top graphs, dashed lines correspo nd to extended \neLLB results (𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅≠0): (a) including thermal fluctuations 𝜉𝜉⃗𝑖𝑖∥,𝜉𝜉⃗𝑖𝑖⊥≠0, and (b) without thermal \nfluctuations ( 𝜉𝜉⃗𝑖𝑖∥=𝜉𝜉⃗𝑖𝑖⊥=0, deterministic case) . In bottom graphs, dashed lines correspond to LLB \nresults (𝜏𝜏⃗𝑖𝑖𝑁𝑁𝑅𝑅=0): (a) including thermal fluctuations 𝜉𝜉⃗𝑖𝑖∥,𝜉𝜉⃗𝑖𝑖⊥≠0, and (b) without thermal \nfluctuations ( 𝜉𝜉⃗𝑖𝑖∥=𝜉𝜉⃗𝑖𝑖⊥=0, deterministic case) . \n \n(c) Except the contrary is indicated, all presented results with the extended eLLB \nmodel were computed with 𝜆𝜆𝑒𝑒𝑒𝑒=0.013. FIG. S5 show the comparison of ASD results \nto eLLB results for three different values of 𝜆𝜆𝑒𝑒𝑒𝑒. A good quanti tative agreement in the \ntime traces of the out -of-plane components of each sublattice predicted by the ASD model \nis achieved time when 𝜆𝜆𝑒𝑒𝑒𝑒=0.013 is considered (central graph in FIG. S5) . For a laser \npulse of 𝜏𝜏 𝐿𝐿=50 fs, and using 𝜆𝜆 𝑒𝑒𝑒𝑒=0.013, the threshold laser power density to achieve \nthe switching is 𝑄𝑄=7.9×1021 W/m3 within the ASD, whereas the threshold is 𝑄𝑄=\n9.9×1021 W/m3 with the eLLB model. However, if 𝜆𝜆𝑒𝑒𝑒𝑒=0.024 both models give the \nsame threshold (𝑄𝑄=9.9×1021 W/m3). \n \n11 \n \nFIG. S5 . Temporal evolution of the out -of-plane component ( 𝑚𝑚𝑧𝑧 vs 𝑡𝑡) calculated with three \ndifferent values of 𝜆𝜆𝑒𝑒𝑒𝑒 as indicated within each graph. Here 𝑄𝑄 =10×10−21 W/m3. The rest of \ninputs are the same as in FIG. S4 . \n \nSN4. Phase diagram in terms of the f luence and the pulse duration \nFIG. 2(a) of the main text presents the phase diagram of the three possible final \nstates as function of the pulse length ( 𝜏𝜏𝐿𝐿) and the absorbed energy from the laser pulse \n(𝑄𝑄). Same results can be also presented in terms of the absorbed fluence, which is given \nby 𝐹𝐹=𝑄𝑄𝜏𝜏𝐿𝐿𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇, where 𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 is the thickness of the ferrimagnetic alloy. This is shown in \nFIG. S 6 for an extended range of laser pulse durations (𝜏𝜏𝐿𝐿). These results are also in \nagreement with recent experimental studies [11] \n \n12 \n \nFIG. S 6. Same r esults as FIG. 2(a) of the main text showing the three possible final states as a \nfunction of the fluence, 𝐹𝐹=𝑄𝑄𝜏𝜏𝐿𝐿𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 for an extended range of pulse durations ( 𝜏𝜏𝐿𝐿). White, red \nand blue colors represent no -switching, deterministic switching (HI-AOS) and thermal \ndemagn etization behaviors respectively . \n \nSN5. Helicity -Dependent AOS (HD -AOS) and Magnetic Circular Dichroism (MCD) \nConsidering the Magnetic Circular Dichroism (MCD) scenario , the absorbed power \nby the FiM under circular polarized laser pulses , 𝑃𝑃(𝑟𝑟,𝑡𝑡), also depends on the magnetic \nstate of the system and the helicity of the pulse . Therefore , 𝑃𝑃(𝑟𝑟,𝑡𝑡) in Eq. (2) of the TTM \nis replaced by 𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁)𝑃𝑃(𝑟𝑟,𝑡𝑡), where 𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁) describes the different absorption \npower for up (𝑚𝑚𝑁𝑁> 0,↑) and down (𝑚𝑚𝑁𝑁< 0,↓) magnetization states as depending on \nthe laser helicity (𝜎𝜎±) of the laser pulse, \n𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁)=�1 +1\n2𝑘𝑘𝜎𝜎±sign(𝑚𝑚𝑁𝑁)� (eS22) \nwhere 𝑚𝑚𝑁𝑁 is the local net out of plane magnetization, ( 𝑚𝑚𝑁𝑁=𝑀𝑀𝑠𝑠𝑇𝑇𝑇𝑇𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇+𝑀𝑀𝑆𝑆𝑅𝑅𝑅𝑅𝑚𝑚𝑧𝑧𝑅𝑅𝑅𝑅), \n𝜎𝜎±= ±1 is the helicity of the laser pulse ( 𝜎𝜎+= +1 for right -handed helicity and 𝜎𝜎−=\n−1 left-handed helicity) , and sign (𝑚𝑚𝑁𝑁) is the sign of the initial net magnetization. Note that \n𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇= ±1 corresponds to 𝑚𝑚𝑧𝑧𝑅𝑅𝑅𝑅=∓1, but 𝑀𝑀𝑠𝑠𝑇𝑇𝑇𝑇 and 𝑀𝑀 𝑠𝑠𝑅𝑅𝑅𝑅 are both positive . 𝑘𝑘 is a f actor \nthat determines the difference in the power absorption with respect to the linearly \npolarized case ( 𝑘𝑘= 0). According to this criterium, a state with local net magnetization \n𝑚𝑚𝑁𝑁> 0 (↑: up) absorb s more energy for a right -handed laser 𝜎𝜎+ and less for left -handed \nhelicity (𝜎𝜎−) than in the linearly polarization case (𝜎𝜎= 0). The opposite happens starting \nfrom a state with local down net magnetization 𝑚𝑚𝑁𝑁< 0 (↓: down). Considering \n𝜓𝜓(𝜎𝜎±,𝑚𝑚𝑁𝑁)𝑃𝑃(𝑟𝑟,𝑡𝑡) in the TTM Eqs (2) -(3) of the main text, we obtain isothermal curves \nof 𝑇𝑇=𝑇𝑇𝑒𝑒=1000 K showing the transition from no-s witching to helicity -dependent \nswitching as a f unction of 𝑄𝑄 and 𝜏𝜏𝐿𝐿 for different combinations of the laser helicity and \nthe initial magnetic state under uniform illumination. \nThe results are shown in FIG. S 7(a) and (b) for 𝑘𝑘= 0.1 and 𝑘𝑘= 0. 02 respectively , \nwhich correspond to differences in the power absorption of 10% and 2% between up and \ndown states . Solid black line corresponds to the case of linear polarization ( 𝜎𝜎= 0) already \n13 \n plotted in Fig. 2(a) of the main text. Cases with (𝜎𝜎+,↑) and (𝜎𝜎−,↓) are represented by the \nsolid -blue curve, whereas solid -red curve corresponds to cases with (𝜎𝜎+,↓) and (𝜎𝜎−,↑). \nAs it can be observed for 𝑘𝑘=0.1 (FIG. S7(a)), the isothermal curves 𝑇𝑇=𝑇𝑇𝑒𝑒=1000 K \nfor combinations (𝜎𝜎+,↑) and (𝜎𝜎−,↓) slightly reduce the values of 𝑄𝑄 and 𝜏𝜏𝐿𝐿 needed to \nachieve switching with respect to the linear polarization case ( 𝜎𝜎=0). On the contrary, \nthe isothermal curve for combinations (𝜎𝜎+,↓) and (𝜎𝜎−,↑) slightly moves towards high \nvalues of 𝑄𝑄 and 𝜏𝜏𝐿𝐿 with respect to linear polarization case ( 𝜎𝜎=0). If 𝑘𝑘 =0.02 (FIG. \nS7(b)), the differences are even smaller, and the three curves almost overlap . This fact \nagrees with experimental observation s, where the Helicity -Dependent AOS is only \nachieved in a narrow range of absorbed power ( 𝑄𝑄) for a fixed pulse length ( 𝜏𝜏𝐿𝐿). Dashed \nlines in FIG. S 7(a)-(b) correspond to the isothermal curves of 𝑇𝑇 =𝑇𝑇𝑒𝑒=1400 K, and \nindicate the transition between s witching and demagnetized m ultidomain behaviors . \n \nFIG. S 7. Phase diagram in the MCD scenario for different values of the MCD coefficient 𝑘𝑘. (a) \nand ( b) show the isothermal curves of 𝑇𝑇=𝑇𝑇𝑒𝑒=1000 K (solid lines) as a function of 𝑄𝑄 and 𝜏𝜏𝐿𝐿 \nindicating the transition between no- switching to switching , as computed from the TTM Eq. (2) -\n(3) for 𝑘𝑘 =0.1, and 𝑘𝑘=0.02 respectively . Dashed curves are isothermal curves indicating the \ntransition between the switching to m ultidom ain regimes ( 𝑇𝑇𝑒𝑒=1400 K). Black curve \ncorresponds to the case of linear polarization (𝜎𝜎=0) already plotted in Fig. 2(a) of the main text. \nCases with (𝜎𝜎+,↑) and (𝜎𝜎−,↓) are represented by the blue curve, whereas red curve corresponds \nto cases with (𝜎𝜎+,↓) and (𝜎𝜎−,↑). Here ↑ and ↓ represent the initial net magnetic state (𝑚𝑚𝑁𝑁), either \nup or down respectively. \n \nSN6. Helicity -Dependent AOS (HD -AOS) and Inverse Fadaray Effect (IFE) \nAs mentioned in the main text, several works claim that the observations of the HD -\nAOS can be ascribed to the Inverse Faraday Effect (IFE) [12] . Within this formalism, \nthe laser pulse generates an effective out -of-plane magneto -optical field whose direct ion \ndepends on the laser pulse helicity as 𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)=𝜎𝜎±𝐵𝐵𝑇𝑇𝑀𝑀𝜂𝜂(𝑟𝑟)𝑛𝑛(𝑡𝑡)𝑢𝑢�⃗𝑧𝑧, where 𝜂𝜂(𝑟𝑟)=\nexp[−4ln(2)𝑟𝑟2/ (2𝑟𝑟0)2 ] is the spatial field profile, and 𝑛𝑛(𝑡𝑡) is its temporal profile. \nNote that there the spatial dependence of 𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) is the same as the one of the absorbed \npower density. A ccording to the literature [12] , the so- called magneto -optical field \n𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) has some persistence with respect to the laser pulse, and therefore its temporal \nprofile is different for 𝑡𝑡<𝑡𝑡0 and 𝑡𝑡>𝑡𝑡0: 𝑛𝑛(𝑡𝑡<𝑡𝑡0)=exp[−4ln(2)(𝑡𝑡−𝑡𝑡0)2/ 𝜏𝜏𝐿𝐿2], and \n14 \n 𝑛𝑛(𝑡𝑡≥𝑡𝑡0)=exp[−4ln(2)(𝑡𝑡−𝑡𝑡0)2/(𝜏𝜏𝐿𝐿+𝜏𝜏𝐷𝐷)2], where 𝜏𝜏𝐷𝐷 is the delay time of the \n𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) with respect to the laser pulse. We have evaluated this scenario by including \nthis field 𝐵𝐵 �⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡) in the effective field of Eq. (4) . The results for the same FiM alloy \nconsidered up to here ( Gdx(FeCo) 1-x, with x =0.25), (see Supplemental Note SN5 ), are \nshown in FIG. S 8 by considering a maximum magneto -optical field of 𝐵𝐵𝑇𝑇𝑀𝑀=20 T with \na delay time of 𝜏𝜏𝐷𝐷=𝜏𝜏𝐿𝐿. \n \n \nFIG. S8. Micromagnetic results of the HD -AOS computed within the IFE scenario. (a) Snapshots \nof the final RE magnetic state ( 𝑚𝑚𝑧𝑧𝑅𝑅𝑅𝑅) after a laser pulse of 𝜏𝜏𝐿𝐿=50 fs for four different values of \nthe absorbed power density ( 𝑄𝑄). Results are shown for four combinations of the initial state ( ↑,↓) \nand helicities ( 𝜎𝜎±) as indicated at the left side. (b) RE magnetic state after every pulse with 𝜎𝜎+ \nfor 𝑄𝑄=5.9×1021 W/m3, showing the appearance of a ring around the central part. The sample \nside is ℓ=20 μm and the laser spot diameter is 𝑑𝑑0=ℓ/2. The IFE was evaluated by considering \na maximum magneto -optical field of 𝐵𝐵 𝑇𝑇𝑀𝑀=20 T with a delay time of 𝜏𝜏𝐷𝐷=𝜏𝜏𝐿𝐿. The material \nparameters correspond to a FiM alloy Gd x(CoFe) 1-x with x =0.25 . \n \nSN7. Inverse Faraday Effect: magneto- optical field or induced magnetic moment \nIn previous section SN6 and in the main text, the Inverse Faraday Effect (IFE) was \nconsidered by adding an effective out -of-plane magneto -optical field (𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)) whose \ndirection depends on the laser pulse helicity . Other authors [13,14] have alternatively \npointed out that under circularly polarized laser pulses , the IFE c an be taken into account \nin micromagnetic simulations by adding a helicity -dependent induced magnetic moment \n15 \n on each sublattice ( Δ𝑚𝑚��⃗𝑖𝑖). We have also explored this alternative by adding, instead of the \n𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡), an additional term to the right -hand side of the eLLB Eq. ( eS7). This term can \nbe expressed as 𝛾𝛾0′ Δ𝑚𝑚��⃗𝑖𝑖, where Δ𝑚𝑚��⃗𝑖𝑖 represents the laser induced magnetization for \nsublattice 𝑖𝑖 :RE,TM, and i t is given by \nΔ𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)=(𝜎𝜎±)𝐾𝐾𝐼𝐼𝐹𝐹𝑅𝑅,𝑖𝑖𝐼𝐼\n𝑐𝑐𝑢𝑢�⃗𝑧𝑧 (eS23) \nwhere 𝐾𝐾𝐼𝐼𝐹𝐹𝑅𝑅,𝑖𝑖 is the IFE constant for sublattice 𝑖𝑖 :RE,TM, 𝐼𝐼=𝑃𝑃(𝑟𝑟,𝑡𝑡)𝑡𝑡𝐹𝐹𝑖𝑖𝑇𝑇 is the laser \nintensity with 𝑃𝑃(𝑟𝑟,𝑡𝑡)=𝑄𝑄𝜂𝜂(𝑟𝑟)𝜉𝜉(𝑡𝑡) (see main text) , and 𝑐𝑐 the speed of light. In FIG. S9 \nwe compare both options to account for the IFE: either by a magneto- optical field \n(𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)≠0 with Δ𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)=0, solid lines in FIG. S9) as described in SN6, or by an \ninduced magnetic moment ( 𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)=0 with Δ𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)≠0, dashed lines in F IG. S9). \nAs it can be observed, both alternatives provide very similar results, so we conclude , as \nin [14] , that both alternatives are essentially equivalent within the scope of our numerical \nstudy . Moreover, the adopted value of 𝜎𝜎±𝐾𝐾𝐼𝐼𝐹𝐹𝑅𝑅,𝑖𝑖=±0.0014 T−1 (see FIG. S9) is also in \ngood quantitative agreement with typical values deduced in [14] from an ab initio \nformalism. \n \nFIG. S9 . Comparison of two micromagnetic implementations of the IFE for both laser helicities: \n(a) 𝜎𝜎+=+1 and (b) 𝜎𝜎−=−1. Solid lines correspond to results obtained assuming that the IFE \ngenerates a magneto -optical field ( 𝐵𝐵�⃗𝑇𝑇𝑀𝑀(𝑟𝑟⃗,𝑡𝑡)≠0), whereas dashed lines are for the induced \nmagnetic moment alternative ( Δ𝑚𝑚��⃗𝑖𝑖(𝑟𝑟⃗,𝑡𝑡)≠0). The values of the maximum magneto -optical field \nis 𝐵𝐵𝑇𝑇𝑀𝑀=20 T, wher eas the IFE constant is 𝜎𝜎±𝐾𝐾𝐼𝐼𝐹𝐹𝑅𝑅 ,𝑖𝑖=±0.0014 T−1. The rest of inputs are the \nsame as in Fig. S8. \n \nSN8. Material inputs for two different compositions \nIn the las t part of the manuscript we studied the probability of switching ( see FIG. \n6) for two FiMs ( Gdx(CoFe) 1-x:RE x(TM) 1-x) with two different compositions : x=0.25 and \nx=0.24. The inputs used in these simulations are collected in the following Table S 1. \n \n x=0.25: Gd x(CoFe) 1-x x=0.24: Gd x(CoFe) 1-x \n𝑀𝑀𝑠𝑠,𝑇𝑇𝑇𝑇 / 𝑀𝑀𝑠𝑠,𝑅𝑅𝑅𝑅 (MA/m) 0.412 / 0.546 0.412 / 0.52 \n𝐴𝐴𝑒𝑒𝑒𝑒,𝑇𝑇𝑇𝑇 / 𝐴𝐴𝑒𝑒𝑒𝑒,𝑅𝑅𝑅𝑅 (pJ/m) 3.27 / 0.189 3.35 / 0 .174 \n𝐽𝐽𝑇𝑇𝑇𝑇−𝑇𝑇𝑇𝑇 / 𝐽𝐽𝑅𝑅𝑅𝑅−𝑅𝑅𝑅𝑅 (×10−21J) 2.59 / 1.35 2.59 / 1.35 \n16 \n 𝐽𝐽𝑇𝑇𝑇𝑇−𝑅𝑅𝑅𝑅 (×1021J) -1.12 \n𝐾𝐾𝑢𝑢,𝑇𝑇𝑇𝑇 / 𝐾𝐾𝑢𝑢,𝑅𝑅𝑅𝑅 (MJ/m3) 1.87 / 0.62 1.89 / 0.59 \n𝜇𝜇𝑇𝑇𝑇𝑇 / 𝜇𝜇𝑅𝑅𝑅𝑅 (𝜇𝜇𝐵𝐵) 1.92 / 7.63 1.92 / 7.63 \n𝛼𝛼𝑇𝑇𝑇𝑇 / 𝛼𝛼𝑅𝑅𝑅𝑅 ( ) 0.02 / 0.02 0.02 / 0.02 \n𝛾𝛾𝑇𝑇𝑇𝑇 / 𝛾𝛾𝑅𝑅𝑅𝑅 (1011×(T⋅s)−1) 1.847 / 1.759 1.847 / 1.759 \n𝑔𝑔𝑇𝑇𝑇𝑇 / 𝑔𝑔𝑅𝑅𝑅𝑅 ( ) 2.1 / 2.0 2.1 / 2.0 \n𝑎𝑎 (nm) 0.352 \n \n𝑘𝑘𝑒𝑒 (W/ (K⋅m)) 91 \n𝐶𝐶𝑒𝑒(300 K) (×105J/ (K⋅m3)) 1.8 \n𝐶𝐶𝑙𝑙 (×106J/ (K⋅m3)) 3.8 \n𝑔𝑔𝑒𝑒𝑙𝑙(300 K) (×1017 W/m3) 7 \n \nTABLE S1. Material inputs adopted to explore the switching probability for two different FiM \nalloys, with different composition s x. These inputs were used to get the results of FIG. 6 in the \nmain text. \n \nSN9. Helicity -Dependent All Optical Switching : MCD & IFE for different \ncompos itions and initial temperatures \nIn FIG. 6 of the main text we explore the switching probability predicted by both \nthe MCD and IFE mechanism s for two different compositions of the FiM and considering \nthat the initial temperature of the thermal bath was room temperature. Similar results can \nbe also obtained by fixing the compos ition of the FiM alloy and chan ging the temperature \nof the thermal bath with a cryostat. These results are shown in FIG. S10 for with x=0.25 \n(left column) and with x =0.24 (right column) compositions and three different \ntemperatures of the thermal bath: 𝑇𝑇 =260 K, 𝑇𝑇=300 K, and 𝑇𝑇 =340 K. For x=0.25, \nboth the IFE and MCD predict similar behavior for 𝑇𝑇 =260 K (FIG. S10(b)) and 𝑇𝑇 =\n300 K (FIG. S10(c) ): respect to the linear polarized laser pulse, the 100% switching \nprobability occurs with smaller Q for circular polarization when (𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇↑) and \n(𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇↓), and with larger Q when (𝜎𝜎+,𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇↑) and (𝜎𝜎−,𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇↓). However, at 𝑇𝑇=\n340 K (FIG. S10(d)), the IFE scenario results in similar behavior but the MCD results are \nthe opposite. This can be easily understood as explained in the main, because the FiM \nalloy with 𝑥𝑥=0.25 is RE -dominated for 𝑇𝑇=260 K and 𝑇𝑇=300 K, whereas becomes \nTM-dominated for 𝑇𝑇=340 K. (Fig. S8(a)). Note that the magnetization compensation \ntemperature for 𝑥𝑥=0.25 is 𝑇𝑇 𝑇𝑇≈320 K. \nSimilar results are also achieved for 𝑥𝑥=0.24 (right column in F IG. S10), but now \nboth IFE and MCD only give similar results for 𝑇𝑇=260 K (FIG. S10(f)), whereas they \npredict opposite behavior for 𝑇𝑇=300 K (FIG. S10(g)) and 𝑇𝑇 =300 K (FIG. S10(h)). \nNote that for 𝑥𝑥=0.24, the FiM is only RE -dominated for temperature below the 𝑇𝑇𝑇𝑇, \nwhich now is 𝑇𝑇 𝑇𝑇≈280 K. \n 17 \n \nFIG. S 10. Temperature dependence of the spontaneous magnetization of each sublattice (RE:Gd; \nTM:CoFe) of the FiM alloy (Gd x(CoFe) 1-x) for two different compositions: (a) x=0.25 and ( e) \nx=0.24. Probability of switching as a function of the absorbed power density ( 𝑄𝑄) for a laser pulse \nof 𝜏𝜏𝐿𝐿=50 fs for different combinations of the initial state ( 𝑚𝑚𝑧𝑧𝑇𝑇𝑇𝑇:(↑,↓)) and the polarization \n(linear: 𝜎𝜎=0 (black dots), and circular 𝜎𝜎±=±1) of the laser pulse as indicated in the legend \nand in the main text: (b) , (c) and (d) corresponds for 𝑇𝑇=260 K, 𝑇𝑇=300 K, and 𝑇𝑇=340 K \nfor 𝑥𝑥=0.25, whereas (f), (g) and (h) to x=0.24. MCD results are shown by solid dots, whereas IFE \nresults are presented by open symbols. Lines are guide to the eyes. \n \n \n18 \n REFERENCES \n[1] U. Atxitia, P. Nieves, and O. Chubykalo -Fesenko, Landau- Lifshitz -Bloch Equation \nfor Ferrimagnetic Materials , Physical Review B 86 , 104414 (2012). \n[2] R. F. L. Evans, W. J. Fan, P. Chureemart, T. A. Ostler, M. O. A. Ellis, and R. W. \nChantrell, Atomistic Spin Model Simulations of Magnetic Nanomaterials , Journal \nof Physics Condensed Matter , 26, 103202 (2014) . \n[3] T. A. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell, U. Atxitia, O. Chubykalo -\nFesenko, S. el Moussaoui, L. le Guyader, E. Mengotti, L. J. Heyderman, F. Nolting, A. Tsukamoto, A. Itoh, D. Afanasiev, B. A. Ivanov, A. M. Kalashnikova, K. \nVahaplar, J. Mentink, A. Kirilyuk, T. Rasing, and A. v. Kimel, Ultrafast Heating as a \nSufficient Stimulus for Magnetization Reversal in a Ferrimagnet , Nature \nCommunications 3, 1666 (2012). \n[4] U. Atxitia, O. Chubykalo -Fesenko, J. Walowski, A. Mann, and M. Münzenberg, \nEvidence for Thermal Mechanisms in Laser -Induced Femtosecond Spin Dynamics , \nPhysical Review B, 81, 174401 (2010). \n[5] E. Beaurepaire, J. -C. Merle, A. Daunois, and J. -Y. Bigot, Ultrafast Spin Dynamics \nin Ferrom agnetic Nickel , Physical Review Letters 76 , 4250 (1996). \n[6] C. Vogler, C. Abert, F. Bruckner, and D. Suess, Stochastic Ferrimagnetic Landau-\nLifshitz -Bloch Equation for Finite Magnetic Structures , Physical Review B 100, \n054401 (2019). \n[7] O. J. Suarez, P. Nieves, D. Laroze, D. Altbir, and O. Chubykalo -Fesenko, Ultrafast \nRelaxation Rates and Reversal Time in Disordered Ferrimagnets , Physical Review \nB. 92, 144425 (2015). \n[8] P. Nieves, U. Atxitia, R. W. Chantrell, and O. Chubykalo -Fesenko, The Classical \nTwo -Sublattice Landau- Lifshitz -Bloch Equation for All Temperatures , Low \nTemperature Physics 41, 739 (2015). \n[9] C. S. Davies, T. Janssen, J. H. Mentink, A. Tsukamoto, A. v. Kimel, A. F. G. van der \nMeer, A. Stupakiewicz, and A. Kirilyuk, Pathways for Single -Shot All -Optical \nSwitching of Magnetization in Ferrimagnets , Physical Review Applied 13, 024064 \n(2020). \n[10] C. T. Ma, X. Li, and S. J. Poon, Micromagnetic Simulation of Ferrimagnetic TbFeCo Films with Exchange Coupled Nanophases , Journal of Magnetism and Magnetic \nMaterials 417, 197 (2016). \n[11] J. Wei, B. Zhang, M. Hehn, W. Zhang, G. Malinowski, Y. Xu, W. Zhao, and S. \nMangin, All-Optical Helicity -Independent Switching State Diagram in Gd - Fe - Co \nAlloys , Physical Review Applied 15 , 054065 (2021). \n[12] K. Vahaplar, A. M. Kalashnikova, A. v. Kimel, S. Gerlach, D. Hinzke, U. Nowak, R. \nChantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing, All -Optical \nMagnetization Reversal by Circularly Polarized Laser Pulses: Experiment and Multiscale Modeling, Physical Review B 85, 104402 (2012). 19 \n [13] M. Battiato, G. Barbalinardo, and P. M. Oppeneer, Quantum Theory of the \nInverse Faraday Effect , Physical Review B. 89, 014413 (2014). \n[14] M. Berritta, R. Mondal, K. Carva, and P. M. Oppeneer, Ab Initio Theory of \nCoherent Laser -Induced Magnetization in Metals , Physical Review Letters 117, \n137203 (2016). \n " }, { "title": "1310.1782v1.Ferrimagnetic_Slater_Insulator_Phase_of_the_Sn_Ge_111__Surface.pdf", "content": "arXiv:1310.1782v1 [cond-mat.mes-hall] 7 Oct 2013FerrimagneticSlater InsulatorPhaseoftheSn/Ge(111)Sur face\nJun-Ho Lee, Hyun-Jung Kim, and Jun-Hyung Cho∗\nDepartment of Physics and Research Institute for Natural Sc iences,\nHanyang University, 17 Haengdang-Dong, Seongdong-Ku, Seo ul 133-791, Korea\n(Dated: June 7, 2021)\nWe have performed the semilocal and hybrid density-functio nal theory (DFT) studies of the Sn/Ge(111)\nsurface to identify the origin of the observed insulating√\n3×√\n3 phase below ∼30 K. Contrasting with the\nsemilocal DFT calculation predicting a metallic 3 ×3 ground state, the hybrid DFT calculation including van\nder Waals interactions shows that the insulating ferrimagn etic structure with√\n3×√\n3 structural symmetry is\nenergetically favored over the metallic 3 ×3 structure. It is revealed that the correction of self-inte raction error\nwith a hybrid exchange-correlation functional gives rise t o a band-gap opening induced by a ferrimagnetic\norder. The results manifest that the observed insulatingph ase is attributedtothe Slatermechanism via itinerant\nmagnetic order ratherthan the hithertoaccepted Mott-Hubb ard mechanism via electron correlations.\nPACS numbers: 71.30.+h, 73.20.At, 75.50.Gg\nTwo-dimensional (2D) electronic systems formed at crys-\ntal surfaceshave attractedmuch attentionbecause of their in-\ntriguing physical phenomena such as charge density waves\n(CDW), magnetic order, Mott insulators, and 2D supercon-\nductivity [1–3]. One of the most popular quasi-2D systems\nis the√\n3×√\n3 phase formed by the 1/3-monolayer adsorp-\ntion of group IV metal atoms, Sn or Pb, on the Si(111) or\nGe(111) surface [4–12]. Here, adatoms locating at T4sites\n[Fig. 1(a)] saturate all the dangling bonds (DBs) of Si or\nGe surface atoms, leaving a single DB for each adatom [13].\nSincetheseDBelectronsareseparatedasfaras ∼7˚A,there-\nsulting narrow half-filled DB band is likely to invoke variou s\ninstabilitiesandstrongelectroncorrelations.\nWe here focus on a prototypical example of quasi-2D sys-\ntems,Sn/Ge(111). Below ∼220K,thissystemundergoesare-\nversible phase transition from a√\n3×√\n3R30◦(hereafter des-\nignated as√\n3×√\n3) structure to a 3 ×3 structure [14]. This\nphase transition was initially interpreted in terms of a sur -\nface CDW formation stabilized by electron correlation ef-\nfects in the√\n3×√\n3 structure [14, 15]. However, the high-\ntemperature√\n3×√\n3 phase was later explained as a dynam-\nical effect in which inequivalent Sn atoms interchange thei r\nverticalpositions[16–18]: i.e.,thethreeSnatoms[up(U) and\ndown (D) atoms in Fig. 1(b)] of two different heights within\nthe3×3structurefluctuatebetweentwopositionsastempera-\nture increases, apparentlyshowinga√\n3×√\n3 structural sym-\nmetry [18]. Nevertheless, the ground state of Sn/Ge(111)\nis still subject to much debate because, as the temperature\nfurther decreases to ∼30 K, various experimental techniques\nsuch as scanning tunneling microscopy (STM), low-energy\nelectron diffraction, and angle-resolved photoemission s pec-\ntroscopy (ARPES) explored a phase transition from the 3 ×3\nstructureto a new√\n3×√\n3 structure wherethree Sn atomsin\n3×3 unit cell have an equivalent height [19]. This structural\nphase transition was observed to accompany simultaneously\na metal-insulator transition (MIT) [19]. However, previou s\ndensity-functional theory (DFT) studies [20–25] with the l o-\ncaldensityapproximation(LDA)andgeneralizedgradienta p-\nproximation(GGA) predicted that the√\n3×√\n3 structure wasnot only metallic but also less stable than the 3 ×3 structure,\nthereby failing to describe the electronic and energetic pr op-\nerties of the observed insulating√\n3×√\n3 phase. To resolve\nthis problem of LDA, Profeta and Tosatti [4] took into ac-\ncountCoulombinteractions(Hubbard U)using the LDA+ U\nscheme. They found that electron correlations can stabiliz e\na magnetic insulator with√\n3×√\n3 structural symmetry,indi-\ncating a Mott-Hubbard insulating ground state. On the other\nhand,Flores et al.pointedoutthattheeffectofelectroncorre-\nlations cannotinduce a transition to a Mott insulating grou nd\nstate [22]. Thus, it remains elusive whether the formation o f\nthe insulating phase of Sn/Ge(111) is attributed to electro n\ncorrelations[26].\n(a) (b)\nD D\n(c)U\nFIG. 1: (Color on line) (a) Top and (b) side views of the metall ic\n3×3 structure of Sn/Ge(111). The side view of the insulating fe rri-\nmagneticstructureisgivenin(c). Thedarkandgraycircles represent\nSn and Ge atoms, respectively. U and D in (b) represent the up a nd\ndownSnatoms,respectively. TheSnatomsarelocatedatthe T4site,\ni.e.,asingle threefoldhollowsiteabove asecondlayerGea tom. For\ndistinction, Ge atoms in the subsurface layers are drawn wit h small\ncircles. In (a), the 3 ×3 and√\n3×√\n3 unit cells are indicated by the\nsolidand dashedlines, respectively. The xandzdirections in(b) are\n[112]and [111], respectively.\nIn the present work, we propose a new origin of the ob-\nserved insulating phase in Sn/Ge(111) due to itinerant mag-\nnetism. This magneticallydriveninsulatingphase via an it in-\nerant single-electron approach is characterized as a Slate r in-\nTypesetbyREVT EX2\nsulator[27]. ItisfoundthattheDBstatesofSnatomsexhibi t\nan itinerant character because of their significant hybridi za-\ntion with the Ge surface states, differing from the case of th e\npreviously proposed Mott-Hubbard insulator [4] where each\nDBelectronwastreatedtobelocalizedatSnadatomsite. We\nnotethat theLDAandGGA tendto stabilizeartificially delo-\ncalized electronic states due to their inherent self-inter action\nerror (SIE), because delocalization reduces the spurious s elf-\nrepulsion of electron [28, 29]. In this regard, previous LDA\nand GGA calculations for Sn/Ge(111) may overestimate the\nstabilityofthemetallic3 ×3structure[20–25]. Itis,therefore,\nvery interesting to examine if the observed insulating phas e\ncan be predicted by the correction of SIE with an exchange-\ncorrelationfunctionalbeyondtheLDAorGGA.\nIn this Letter, we present a new theoretical study for\nSn/Ge(111) based on the hybrid DFT scheme including\nvan der Waals (vdW) [30] interactions (termed DFT+vdW\nscheme). We find that the correction of SIE with the hybrid\nexchange-correlation functional of Heyd-Scuseria-Ernze rhof\n(HSE) [31] stabilizes the insulating ferrimagnetic (FI) st ruc-\nture where the band-gap opening occurs by a FI spin order-\ning of three Sn atoms within the 3 ×3 unit cell. Here, the\nbuckling of three Sn atoms is suppressed to show apparently\na√\n3×√\n3 structural symmetry, and the stability of the FI\nstructure is further enhanced by the inclusion of vdW inter-\nactions. The calculated magnetic moment of the FI structure\nis well distributed over Sn adatoms as well as Ge substrate\natoms, giving rise to a magnitude of 1 µBper 3×3 unit cell.\nIt is thus demonstrated that the observed insulating phase i n\nSn/Ge(111) can be represented as a Slater insulator through\nitinerantmagnetism,notasthepreviously[4]proposedMot t-\nHubbardinsulatorbyCoulombinteractions U.\nThe present semilocal and hybrid DFT calculations were\nperformed using the FHI-aims [32] code for an accurate,\nall-electron description based on numeric atom-centered o r-\nbitals, with “tight” computationalsettings. For the excha nge-\ncorrelation energy, we employed the GGA functional of\nPerdew-Burke-Ernzerhof (PBE) [33] as well as the hybrid\nfunctional of HSE [31]. The k-space integration was done\nwith the 15 ×15 and 9×9 uniformmeshesin the surface Bril-\nlouin zones of the√\n3×√\n3 and 3×3 unit cells, respectively.\nThe Ge(111) substrate was modeled by a 6-layer slab with\n∼34˚Aofvacuuminbetweentheslabs[34]. Here,weusedthe\noptimizedGelatticeconstants a0=5.783,5.718,and5.667 ˚A\nfor the PBE, HSE, and HSE+vdW calculations, respectively.\nThe HSE+vdWlattice constant [35, 36] agreesmost with the\nexperimentalvalueof5.658 ˚A[37]. EachGe atominthebot-\ntom layer was passivated by one H atom. All atoms except\nthe bottom layer were allowed to relax along the calculated\nforces until all the residual force components were less tha n\n0.02 eV/ ˚A. The employed HSE+vdW scheme was success-\nfully applied to determine the energy stability of the metal lic\nandinsulatingphasesinindiumnanowiresonSi(111)[38].\nWe begin to optimize the nonmagnetic (NM)√\n3×√\n3 and\n3×3 structures using the PBE functional. The optimized top\nand side views of the 3 ×3 structure are displayed in Fig.TABLEI:Calculatedtotalenergies(inmeVper√\n3×√\n3unitcell)of\nthe 1U2D, FM, and FI structures relative to the NM√\n3×√\n3 struc-\nture. For comparison, the previous LDA and GGA results are al so\ngiven.\n1U2D FM FI\nPBE −10.1 − −\nHSE+vdW −31.6 −18.7 −37.7\nLDA (Ref. [21]) −5 − −\nLDA (Ref. [24]) −7.5 − −\nLDA (Ref. [4]) −9 − −\nGGA (Ref. [17]) −5 − −\n-1-0.5 0 0.5 1\nΓ KM ΓEnergy (eV)\n-1-0.5 0 0.5 1\nΓ KM ΓEnergy (eV)\nS1S1S2ΓΜ\nΚ\nS3\n→S2S3\n→\nDUDDUD\n(b)(a)→\nFIG.2: (Coloronline)(a)Surfacebandstructureofthe1U2D struc-\nture computed using the PBE functional. The spin-polarized surface\nband structure of the FI structure computed using the HSE+vd W\nscheme is given in (b). The band dispersions are plotted alon g the\nsymmetrylinesoftheBrillouinzone ofthe3 ×3unitcell[seethein-\nset in(a)]. The Γ-M line corresponds to [11 2] direction. The charge\ncharacters of the DB states at the Γpoint are also displayed with an\nisosurfaceof0.006e/ ˚A3. TheenergyzerorepresentstheFermilevel.\nThe majority and minoritybands in (b) are drawn with the soli d and\ndashed lines,respectively.\n1(a) and 1(b), respectively. It is seen that the U atom posi-\ntions higher than the two D atoms by 0.35 ˚A, in good agree-\nment with those (ranging from 0.26 to 0.36 ˚A) of previous\nDFT studies [4, 20–25, 39]. This so-called 1U2D structure\nis more stable than the NM√\n3×√\n3 structure by 10.1 meV\nper√\n3×√\n3 unit cell, which is well comparable with previ-\nous DFT results (see Table I). As shown in Fig. 1S of the\nSupplemental Material [40] [Fig. 2(a)], the calculated ban d\nstructure for the NM√\n3×√\n3 (1U2D) structure exhibits the3\npresenceofoccupiedDBstate(s)attheFermilevel,indicat ing\na metallic feature. It is revealed that the charge character s of\nthe DB states in the 1U2D structure [see Fig. 2(a)] represent\na chargetransferfromtheD to the U atoms[24]. Thischarge\ntransfergivesrisetoareducedCoulombrepulsionbetweenS n\nDB electrons, resulting in a more stabilization over the NM√\n3×√\n3 structure. Ourspin-polarizedPBE calculationswere\nnotabletofindanyspinorderingwithinthe√\n3×√\n3and3×3\nunit cells. Thus, we can say that the semilocal DFT scheme\nwiththePBEfunctionalcannotpredicttheinsulating√\n3×√\n3\nphaseobservedbelow ∼30K [19].\nItisnoteworthythatarecenthigh-resolutionphotoemissi on\nstudy [5] for the Sn 4 dcore level of the metallic 3 ×3 struc-\nture resolved three components,which were assigned to each\nof the three Sn atoms within the 3 ×3 unit cell. On the basis\nofthisphotoemissiondatatogetherwith STMimages,Tejeda\net al.[5] concluded that the two D atoms position at slightly\ndifferentheights, formingan inequivalent-down-atoms(I DA)\nstructure. Our PBE calculation shows that the IDA struc-\nture with a height difference of ∼0.03˚A between the two D\natoms is almost degenerate in energy (less than 0.1 meV per\nSn atom)with the 1U2D structure,consistent with a previous\nLDA calculation [39]. For the 1U2D and IDA structures, we\ncalculate the Sn 4 dcore-levelshifts using initial-state theory,\nwhere the shift is defined by the difference of the eigenval-\nues of the Sn 4 dcore level at different sites. The results are\ndisplayedinFig. 3andcomparedtothephotoemissionexper-\niment[5]. WefindthateachSn4 dcorelevelissplitintothree\nsublevels, i.e., the degenerate Cd1(Cd2) sublevel arising from\nthedxyanddx2−y2(dyzanddxz) orbitals and the Cd3sublevel\nfrom the dz2orbital. Note that such a crystal field splitting\nis conspicuous for the two D atoms but negligible for the U\natom. On the basis of the calculated initial-state core leve ls,\nthe observed C1,C2, andC3components (see Fig. 3) can be\nassociated with the three sublevels of the U atom, Cd1of the\ntwo D atoms, and Cd2andCd3of the two D atoms, respec-\ntively. Thus, we can say that the observed two components\nof higher binding energy are attributed to the effect of crys -\ntal field splitting [41] on the Sn 4 dcore levels of the two D\natoms,ratherthanto differentcorelevelsforthetwo inequ iv-\nalentDatomsintheIDAstructure[5]. Theinitial-statethe ory\nfor the 1U2D (IDA) structure shows that the Cd1,Cd2, and\nCd3sublevels for the two D atoms shift to higher binding en-\nergy by 155 (157 ±1), 237 (238 ±2), and 256 (257 ±3) meV,\nrespectively, relative to the average value of three sublev els\nfor the up atom. These shifts are smaller than those (230 ±40\nand 390±40 meV) of C2andC3relative to C1, which were\nresolved from the high-resolution photoemission spectra [ 5].\nThis differenceof Sn 4 dcore-level shifts between the initial-\nstatetheoryandthephotoemissionexperimentmayreflectth e\nfinal-statescreeningeffects[42].\nIn order to provide an explanation for the observed insu-\nlating√\n3×√\n3 phase, Profeta and Tosatti [4] performed the\nLDA +Ucalculation to propose the Mott-Hubbardinsulator,\nwhere the inclusion of electron correlations strongly modi -\nfies the ground state of Sn/Ge(111) from a NM 3 ×3 metal0.40.30.20.1 00.40.30.20.1 0Experiment\nC3 C2 C1\nTheory-1U2Dz z\nd2z\ndC3dC2dC1 U\nD1\nD2\n0.40.30.20.1 0Theory-IDA\nBinding energy (eV) dC3dC2dC1\ndyz\nd2x2-y dxzdxy\nx y x y\nFIG. 3: (Color on line) Calculated Sn 4 dsurface core-level shifts of\nthe1U2DandIDAstructures,incomparisonwiththehigh-res olution\nphotoemission experiment [5]. The shifts for the two D atoms (D1\nand D2) are given with respect to the average of the three subl evels\nfor the U atom. The positive sign indicates a shift to higher b inding\nenergy. The five dorbitals of the D1 atom are displayed with an\nisosurface of ±0.2(e/˚A3)1/2.\ntoamagneticinsulatorwith√\n3×√\n3structuralsymmetry[4].\nHere, the FI spin order in the Mott-Hubbard insulating state\nwasstabilizedwithina localizedorHeisenbergpicturewhe re\nanunpairedelectronwaslocalizedateachSnadatomsitewit h\na spin moment of 1 µB. This localized picture for magnetism\ncontrastswiththepresentitinerantorSlatermagnetismwh ere\nthe magnetic moment is well delocalized over Sn adatoms\nas well as Ge substrate atoms, as discussed below. Since\nthe Sn DB state is largely delocalized through hybridizatio n\nwith the Ge surface states [see Fig. 2(a)], the SIE inherent\nto the PBE functional may cause the incorrect prediction of\nthe metallic 3 ×3 structure as a ground state. To circumvent\nsuch an over-delocalization of Sn DB electrons, we use the\nHSE+vdW scheme to optimize the NM and magnetic struc-\ntures of Sn/Ge(111). The calculated total energies of the\n1U2D, FI, and ferromagnetic (FM) structures relative to the\nNM-√\n3×√\n3 structure are given in Table I. We find that the\nFI structure is energetically favored over the 1U2D and FM\nonesby6.1and19.0meVper√\n3×√\n3unitcell,respectively.\nThe optimized geometry of the FI structure is shown in Fig.\n1(b), where the buckling of three Sn atoms within the 3 ×3\nunit cell is suppressed to become flat, leading to a√\n3×√\n3\nstructural symmetry [43]. In Fig. 2(b), the calculated band\nstructure of the FI structure shows a band-gap opening of 71\nmeV, in goodagreementwith the ARPES measurementof60\nmeV [19]. Therefore, the present HSE+vdW calculation pre-4\ndicts an insulating FI ground state with√\n3×√\n3 structural\nsymmetry, consistent with the observed insulating√\n3×√\n3\nphase[19].\nSince the total energyobtained from the HSE+vdW calcu-\nlation is composed of the HSE energy ( EHSE) and the vdW\nenergy(EvdW), the total energydifferencebetween the 1U2D\nand FI structures can be divided into the two components,\nΔEHSEandΔEvdW. For this decomposition,we obtain ΔEHSE\n=4.9meV[44],whichislargerthan ΔEvdW=1.2meV.Thus,\nwecansaythatthecorrectionofSIEwiththeHSEfunctional\ngivesa more dominantcontributionto the stabilization of t he\ninsulatingFIstructureoverthemetallic1U2Dstructure,c om-\nparedtothat fromvdWinteractions.\nFigure2(b)alsoshowsthechargecharactersofthespin-up\n(denotedas S1↑andS3↑)andspin-down( S2↓)DBstatesinthe\nFIstructure. Itisrevealedthat S1↑andS3↑representsomehy-\nbridization of two DB electrons. All of the three DB states\nstronglyhybridizewith the pzorbitalsofthesurfaceandsub-\nsurface Ge atoms. This strong hybridization gives rise to a\nlarge delocalization of spin moments up to deeper Ge atomic\nlayers (see Fig. 4). The sum ( m) of the spin moments of\nSn atoms or Ge atoms in each layer is also given in Fig. 4.\nHere, the spin moment of each atom is calculated by Mul-\nliken analysis. We find that the Sn layer has m= 0.22µB,\nwhile the first and third Ge layers have m= 0.40 and 0.22\nµB, respectively, which are significantly larger than those ob -\ntained from other Ge layers. Note that the total spin moment\nis1µBper3×3unitcell. Theresultofalargespindelocaliza-\ntion over Sn atoms and Ge substrate atoms contrasts with the\ncase of the previously proposed Mott-Hubbard insulator [4]\nwhereeachSn atom hasa localizedspinmomentof1 µBasa\nconsequenceofelectroncorrelations. Itisremarkabletha tthe\npresent FI order is determined by an itinerant single-elect ron\napproach with the correction of SIE, thereby representing a\nSlaterinsulatordrivenbyitinerantmagnetism.\n(µ )Bm\nSn layer 0.22\n1st Ge layer 0.40\n2nd Ge layer 0.03\n3rd Ge layer 0.22\n4th Ge layer 0.04\n5th Ge layer 0.05\n6th Ge layer 0.03\nFIG.4: (Coloronline)SpindensityoftheFIstructure. Them ajority\n(minority)spindensityisdisplayedindark(bright)color withaniso-\nsurfaceof0.01( −0.01)e/˚A3. Thesum( m)ofthespinmomentsofSn\natoms or Ge atoms in each layer is also given. For the spin mome nt\nof eachGe atom, see Table ISof the Supplemental Material[40 ].\nWe note that the total energy difference ΔEbetween the\nmetallic 1U2D structure and the insulating FI structure is\n6.1 meV per Sn atom. Although the MIT temperature can\nbe predicted by the precise entropy-related free energy dif -\nference (TΔS) between the 1U2D and the FI structures, we\nroughly estimate it by considering only the electronic con-tribution to the entropy, as done by Profeta and Tosatti [4].\nAssuming that the FI structure has a lack of spin entropy\nand a charge gap, TΔSwas approximated to γT2whereγis\nthe electronicspecific heat coefficient(roughlyof order ∼0.1\nmeV/site K2) [4, 45]. We thus estimate the MIT temperature\nas∼8K,whichissomewhatbelowtheobservedMITtemper-\nature of∼30 K [19]. This deviation of the MIT temperature\nmayreflect that the 1U2Dand FI structureshavedifferentvi-\nbrationalcontributionstotheentropy,whicharenottaken into\naccountin TΔS.\nIn conclusion, our semilocal and hybrid DFT calculations\nshowed the different predictions for the ground state of the\nSn/Ge(111) surface. Contrasting with the PBE functional\npredicting a metallic 3 ×3 ground state, the HSE functional\nshowed that the correction of SIE cures the delocalization\nerror to predict an insulating FI ground state with√\n3×√\n3\nstructural symmetry. We found that the magnetic moment of\nthe FI structure is well distributed over Sn adatoms as well\nas Ge substrate atoms. It is thus demonstrated that the ob-\nserved insulating phase in Sn/Ge(111) can be represented as\na Slater insulator through itinerant magnetism rather than a\nMott-Hubbardinsulator driven by Coulomb interactions. We\nnotice that the Sn/Si(111) surface has also been much stud-\nied to determine its exact crystallographicarrangement,e lec-\ntronic structure, and ground state [9, 10]. Similar to the\npresent case of Sn/Ge(111), we anticipate that the correcti on\nofSIEwouldbeofimportancetodescribethestructural,ele c-\ntronic,andenergeticpropertiesoftheisoelectronicSn/S i(111)\nsystem.\nThis work was supported by National Research Founda-\ntionofKorea(NRF)grantfundedbytheKoreanGovernment\n(NRF-2011-0015754). The calculations were performed by\nKISTI supercomputing center through the strategic support\nprogram(KSC-2012-C3-18)for the supercomputingapplica-\ntionresearch.\n∗Correspondingauthor: chojh@hanyang.ac.kr\n[1] E. Tosatti, in Electronic Surface and Interface States on\nMetallic Systems , edited by E. Bertel and M. Donath (World\nScientific,Singapore, 1995), p.67.\n[2] J. M. Carpinelli, H. H. Weitering, E. W. Plummer, and R.\nStumpf, Nature 381, 398 (1996).\n[3] H.H.Weitering,J.M.Carpinelli,A.V.Melechko,J.Zhan g,M.\nBartkowiak, E.W.Plummer,Science 285, 2107 (1999).\n[4] G. Profetaand E.Tosatti,Phys.Rev. Lett. 98, 086401 (2007).\n[5] A. Tejeda, R. Cort´ es, J. Lobo-Checa, C. Didiot, B. Kierr en, D.\nMalterre, E. G. Michel, and A. Mascaraque, Phys. Rev. Lett.\n100, 026103 (2008).\n[6] S. Colonna, F. Ronci, A. Cricenti, and G. Le Lay, Phys. Rev .\nLett.101, 186102 (2008).\n[7] H. Morikawa and H. W. Yeom, Phys. Rev. Lett. 102, 159601\n(2009).\n[8] S. Colonna, F. Ronci, A. Cricenti, and G. Le Lay, Phys. Rev .\nLett.102, 159602 (2009).\n[9] G. Li, P. H¨ opfner, J. Sch¨ afer, C. Blumenstein, S. Meyer , A.5\nBostwick,E.Rotenberg,R.Claessen,andW.Hanke,Nat.Com-\nmun.4, 1620 (2013), and references therein.\n[10] P.Hansmann, T.Ayral,L.Vaugier,P.Werner,andS.Bier mann,\nPhys.Rev. Lett. 110, 166401 (2013).\n[11] S. Modesti, L. Petaccia, G. Ceballos, I. Vobornik, G. Pa nac-\ncione, G. Rossi, L. Ottaviano, R. Larciprete, S. Lizzit, and A.\nGoldoni, Phys.Rev. Lett. 98, 126401 (2007).\n[12] F. Ronci, S. Colonna, A. Cricenti, and G. Le Lay, Phys. Re v.\nLett.99, 166103 (2007).\n[13] J.E.Northrup, Phys. Rev. Lett. 53, 683(1984).\n[14] J. M. Carpinelli, H. H. Weitering, M. Bartkowiak, R. Stu mpf,\nandE.W.Plummer, Phys.Rev. Lett. 79, 2859 (1997).\n[15] A.Goldoni andS.Modesti, Phys. Rev. Lett. 79, 3266 (1997).\n[16] R. I. G. Uhrberg and T. Balasubramanian, Phys. Rev. Lett .81,\n2108 (1998).\n[17] J. Avila, A. Mascaraque, E. G. Michel, M. C. Asensio, G. L e\nLay,J. Ortega,R. P´ erez,and F.Flores,Phys.Rev. Lett. 82, 442\n(1999).\n[18] F.Ronci,S.Colonna, S.D.Thorpe, A.Cricenti,andG.Le Lay,\nPhys.Rev. Lett. 95, 156101 (2005).\n[19] R. Cort´ es, A. Tejeda, J. Lobo, C. Didiot, B. Kierren, D. Mal-\nterre, E. G. Michel, and A. Mascaraque, Phys. Rev. Lett. 96,\n126103 (2006).\n[20] S. de Gironcoli, S. Scandolo, G. Ballabio, G. Santoro, a nd E.\nTosatti,Surf.Sci. 454, 172 (2000).\n[21] J. Ortega,R. P´ erez, and F. Flores,J. Phys. Condens. Ma tter12,\nL21(2000).\n[22] F. Flores, J. Ortega, R. P´ erez, A. Charrier, F. Thibaud au, J. M.\nDebever, and J.M. Themlin, Prog.Surf.Sci. 67, 299 (2001).\n[23] R. P´ erez, J. Ortega, and F. Flores, Phys. Rev. Lett. 86, 4891\n(2001).\n[24] O. Pulci, M. Marsili, P. Gori, M. Palummo, A. Cricenti, F .\nBechstedt, andR. DelSole,Appl. Phys.A 85, 361 (2006).\n[25] P. Gori, F. Ronci, S. Colonna, A. Cricenti, O. Pulci, and G. Le\nLay,Europhys. Lett. 85, 66001 (2009).\n[26] A. Tejeda, Y. Fagot-R´ evurat, R. Cort´ es, D. Malterre, E. G.\nMichel, and A. Mascaraque, Phys. Status Solidi A 209, 614\n(2012), and references therein.\n[27] J.C.Slater,Phys.Rev. 82, 538 (1951).\n[28] J.P.Perdew andA. Zunger, Phys.Rev. B 23, 5048 (1981).\n[29] S.K¨ ummel andL.Kronik, Rev. Mod. Phys. 80, 3(2008).\n[30] A. Tkatchenko and M. Scheffler, Phys. Rev. Lett. 102, 073005\n(2009).\n[31] A.V.Krukau,O.A.Vydrov,A.F.Izmaylov,andG.E.Scuse ria,\nJ.Chem. Phys. 125, 224106 (2006).[32] V. Blum, R. Gehrke, F. Hanke, P. Havu, V. Havu, X. Ren, K.\nReuter, and M. Scheffler, Comput. Phys. Commun. 180, 2175\n(2009).\n[33] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett .77,\n3865 (1996); 78, 1396(E) (1997).\n[34] Calculations withthek-pointsamplingof20 ×20meshesinthe\nsurface Brillouin zone of the√\n3×√\n3 unit cell, the eight-layer\nslab of Ge(111), and the force criterion of 0.01 eV/ ˚A change\nthe energy difference between the 1U2D and FI structures by\nlessthan0.2meVper√\n3×√\n3unitcell.Itisalsofoundthatthe\ncalculatedspinmomentsofSnandGeatomsintheFIstructure\nchange littleby less than0.01 µB.\n[35] G.-X.Zhang, A.Tkatchenko, J.Paier,H.Appel,and M.Sc hef-\nfler,Phys. Rev. Lett. 107, 245501 (2011).\n[36] N. Marom, A. Tkatchenko, M. Rossi, V. V. Gobre, O. Hod, M.\nScheffler, and L. Kronik, J. Chem. Theory Comput. 7, 3944\n(2011).\n[37] C.Kittel,in IntroductiontoSolidStatePhysics ,8thed.(Wiley,\nNew York, 2005), p.20.\n[38] H.-J.Kim and J.-H.Cho, Phys. Rev. Lett. 110, 116801 (2013).\n[39] S.¨Ozkaya, M. C ¸akmak, and B. Alkan, Phys. Status Solidi B\n248, 2142 (2011).\n[40] See Supplemental Material at\nhttp://link.aps.org/supplemental/xxxx for the band stru c-\nture of the√\n3×√\n3 structure computed using the PBE\nfunctional and the spinmoments of Snand Geatoms.\n[41] F.A.Cotton, Chemical Applicationsof GroupTheory (Wiley,\nNew York, 1990).\n[42] E.Pehlke andM. Scheffler,Phys. Rev. Lett. 71, 2338 (1993).\n[43] Since the HSE calculation without vdW interactions pre dicts\nthe insulating FI ground state with the disappearance of Sn\nbuckling, we can say that the flat structure of Sn atoms is not\ndue to vdW interactions but tothe FIspin ordering between Sn\natoms. It is found that the Sn-Ge bond lengths and the height\ndifference between the up and down Sn atoms (in the 1U2D\nstructure) obtained usingHSE+vdWdecrease bylessthan0.0 2\n˚A, compared tothose obtained using HSE.\n[44] It is notable that ΔEHSEobtained using the HSE calcula-\ntion without vdW interactions is 2.0 meV. This HSE value\nis somewhat different from that (4.9 meV) obtained from the\nHSE+vdW calculation, possibly due to the use of two differen t\nGe lattice constants optimized usingHSEand HSE+vdW.\n[45] K. Kanoda, J.Phys.Soc. Jpn. 75, 051007 (2006)." }, { "title": "1907.00647v1.Robust_Formation_of_Ultrasmall_Room_Temperature_Neél_Skyrmions_in_Amorphous_Ferrimagnets_from_Atomistic_Simulations.pdf", "content": "1 \n Robust Formation of Ultrasmall Room -Temperature Neél \nSkyrmions in Amorphous Ferrimagnets from Atomistic \nSimulations \nChung Ting Ma1, Yunkun Xie2, Howard Sheng3, Avik W. Ghosh1,2, and S. Joseph Poon1* \n1Department of Physics , University of Virginia, Charlottesville, Virginia 22904 USA \n2Department of Electrical and Computer Engineering , University of Virginia, Charlottesville, \nVirginia 22904 USA \n3Department of Physics and Astronomy, George Mason University, Fairfax, Virginia 22030 USA \n* sjp9x@virginia.edu \n \nNeél skyrmions originate from interfacial Dzyaloshinskii Moriya interacti on (DMI) . Recent studies \nhave explored us ing thin -film ferromagnets and ferrimag nets to host Ne él skyrmions f or \nspintronic applications. However, it is unclear if ultrasmall (10 nm or less) skyrmions can ever be \nstabilized at room temperature for practical use in high density parallel racetrack memories. \nWhile thicker films can improve stability, DMI decay s rapid ly away from the interface. As such, \nspins far away from the interface would experience near -zero DMI, raising question on whether \nor not unrealistically large DMI is needed to stabilize skyrmions, and whether skyrmions will also \ncollapse away from the in terface. To address these questions, we have employed atomistic \nstochastic Landau -Lifshitz -Gilbert simu lations to investigate skyrmion s in amorphous \nferrimagnetic GdCo . It is revealed that a significant reduction in DMI below that of Pt is sufficient \nto st abilize ultrasmall skyrmions even in films as thick as 15 nm . Moreover, skyrmions are found \nto retain a uniform columnar shape across the film thickness due to the long ferrimagnetic \nexchange len gth despite the decaying DMI . Our r esults show that increasing thickness and \nreducing DMI in GdCo can further reduce the size of skyrmions at room temperature, which is \ncrucial to improve the density and energy efficiency in skyrmion based devices. \nIntroduction \nMagnetic skyrmions have topologically protected spin textures. Their potential in advancing \nmemo ry density and efficiency has drawn extensive investigation in recent years1-25. In magnetic \nmaterials, skyrmions are stabilized th rough the Dzyaloshi nskii Moriya interaction (DMI)26-27, \ngenerated by either inherent chiral asymmetries or by interfacial symmetry breaking . Intrinsic \nDMI arises in non-centrosymmetric crystal s such as B20 alloys , where Bloch skyrmions have been \nfound in MnSi and FeGe at low temperature12-13. On the other hand , interfacial DMI originates \nfrom inversion symmetry breaking by an interfacial layer with strong spin-orbit coupling . \nMultilayer stacks, such as Pt/Co/Os/Pt, Ir/Fe/Co/Pt and Pt/Co/Ta , have been found to host 40 nm \nto over 1 μm Ne él skyrmions at room temperature14-16. Several challenges remain in developing \nskyrmion based memory and logic devices - for instance , skyrmion Hall effect can pr esent a \nsignificant challenge in guid ing skyrmion s linearly along racetracks 19-22. More critically , aggressive \nreduction in skyrmion sizes is needed to optimize skyrmion based devices , whereupon their room \ntemperature stability becomes a problem . Thicker magnetic layers are required in most cases to \nincrease stability17-18. However , for ferromagnet (FM) /heavy meta l multilayer stacks, increase in \nthickness of the magnetic layer can lead to a loss of interfacial anisotropy and the reduction of \nthe strength of average DMI28-31, both of which are critical for skyrmion formation. To overcome \nthese challenges, we need to consider a suite of materials and understand their underlying physics , \nespecially with varying film thickness. 2 \n Amorphous rare -earth-transitional -metal ( RE-TM) ferrimagnet s (FiM) are potential candidates to \novercome these challenges. Several properties of RE -TM alloys provide a favorable environment \nto host small skyrmion s at room temperature. Their isotropic structure helps with avoiding defect \npinning18, while their intrinsic perpendicular magnetic anisotropy (PMA) 32-35 helps stabil ize small \nskyrmion s by allowing the use of thicker films ( > 5 nm) . However, the effectiveness of interfacial \nDMI decreases significantly away from the interface28-31, which is the focus of our present \ninvestigation. Besides PMA, the magnetization of RE -TM alloys vanishes at the compensation \ntemperature36. With near zero magnetization and near compensated angular momentum37, the \nskyrmion Hall effect is vastly reduced18-20,23, and the skyrmion velocity is predicted to be maximum \nnear the compensation point of angular moment um61. Recently, all-optical helicity dependent \nultrafast switching has been demonstrated in RE-TM alloys using a circularly polarized laser38-45. \nThis gives an additional tool to control spins in device structures . Indeed, RE-TM alloys have begun \nto draw interest in the field of skyrmion research . Large skyrmions of ~150 nm have been \nobserved in Pt/Gd FeCo/MgO23, and skyrmion bound pairs have been found in Gd/ Fe multilayers24. \nRecently , small skyrmions approaching 10 nm were found in Pt/GdCo/TaO x films25. Understanding \nthese results will enable f urther tuning to reduce the size of the skyrmion s. To guide experiments , \nnumerical simulation has served as an important tool , especially for co mplex systems such as RE-\nTM alloys40,46-50. Several methods, such as atomistic Landau -Lifshitz -Gilbert (LLG) algorithm40,46-49 \nand micromagnetic Landau -Lifshitz -Bloch (LLB) algorithm49 have been employed to provide \ndeeper understanding of the magnetic properties in RE-TM alloys . \nIn this study , an atomistic LLG algorithm40,46-48 is employed to investigate the properties of \nskyrmions in Gd Co with interfacial DMI . Although the sign of DM I at FM/heavy metal interface is \nwell studied51-57, the sign of DMI involv ing a FiM remains complex and is rarely discussed . Here, \nwe consider two scenario s for the DMI between Gd and Co (dGd-Co). First, DMI between the \nantiferromagnetic (AFM) pair is set to the same sign as DMI between ferromagnetic pair, i.e. dGd-\nCo > 0. Second , the case of d Gd-Co < 0 is considered. The latter appears to be f avored by the sign of \nAFM inter action27. Moreover, t o incorporate DMI being an interfacial effect, an exponential ly \ndecay ing DMI is utilized. Simulation results find that with a switched DMI sign, near 10 -nm \nskyrmions remain robust in Gd Co films as thick as 15 nm at room temperature . Through numerical \ntomography maps , we find that skyrmions at room temperature are distributed as a near uniform \ncolumn in thicker samples , despite a spatially decaying DMI. \nResults and Discussion \nWe will now begin to investigate if ultrasmall skyrmions in GdCo can survive an exponential DMI \nreduction over thick sample sizes. To incorporate the amorphous nature of GdCo, we employ an \namorphous structure of RE 25TM 75 from ab initio molecular dynamics calculations , as shown in Fig. \n1. As shown in Fig. 2 , at 300 K , the magnetization of amorphous Gd 25Co75 is 5 x 104 A/m , and it has \na compensation temperature near 250 K. We begin with a n exponential DMI decay away from the \ninterface, as shown in Fig. 3. The DMI value discussed herein is the interfacial DMI D 0. The decay \nlength of DMI is based on both previous simulations and experiments. DMI calculation in Co/Pt \ninterface finds a significant decrease in DMI beyond the second layer of Co from the Pt interface30. \nExperimental results also find similar decay in Co -Alloy/Pt interface52-54. In amorphous GdCo, we \nadapted the “second layer” as decay length for direct comparison wi th these findings. \nA range of interfacial DMI values, from dCo-Co = 0.1 x 10-22 J to dCo-Co = 2.0 x 10-22 J (D = 0.12 to 2. 38 \nmJ/m2), and three thicknesses, 5 nm, 10 nm, and 15 nm, are considered. Only those show \nskyrmions are shown herein. To shorten computational time, thicker samples of 10 nm and 15 nm 3 \n are simulated using a 5 nm thick sample by conserving DMI energy density across the film. To \nconserve the total DMI energy, a faster decay is employed in a 5 nm sample to simulate 10 nm \nand 15 nm thick samples to keep the sum of DMI energy to be the same. To check the validity of \nthis simplification, we have compared the results of 10 nm thick samples and 5 nm thick samples \nwith faster DMI decay to verif y that the two sets of samples produce identical results. First, we \nconsider two scenario s for the sign of d Gd-Co, as both + and - signs have been reported in \nantiferromagnetically coupled systems59,60. Fig. 4 shows the color maps of equilibrium spin \nconfigurations at 300 K for both d Gd-Co > 0 and d Gd-Co < 0 . For the case of d Gd-Gd, dCo-Co > 0 and dGd-\nCo > 0, the simulation with dCo-Co = 0.25 x 10-22 J, dGd-Gd = 2.96 x 10-22 J and dGd-Co = 0.86 x 10-22 J \ncorresponds to an average DMI of D = 0. 21 mJ/m2. The value of dGd-Gd and dGd-Co is calculated from \ndCo-Co by multiplying the ratio of Gd moment μGd over Co moment μCo. Further discussion in the \nsupplementary material shows that for a given average DMI, the energy minimum and thus \nskrymion size is independent of how each DMI term varies . Eq. 3 shows the formula used for \nconverting atomistic DMI to average DMI for Gd xCo1-x. \nD=2\nπ1\n𝑛̅[(1−𝑥)(𝑛𝐶𝑜−𝐶𝑜 ̅̅̅̅̅̅̅̅̅𝑑𝐶𝑜−𝐶𝑜\n𝑟𝐶𝑜−𝐶𝑜 ̅̅̅̅̅̅̅̅2+𝑛𝐶𝑜−𝐺𝑑 ̅̅̅̅̅̅̅̅̅|𝑑𝐺𝑑−𝐶𝑜|\n𝑟𝐶𝑜−𝐺𝑑 ̅̅̅̅̅̅̅̅2)+x(𝑛𝐺𝑑−𝐺𝑑 ̅̅̅̅̅̅̅̅̅𝑑𝐺𝑑−𝐺𝑑\n𝑟𝐺𝑑−𝐺𝑑 ̅̅̅̅̅̅̅̅̅2+𝑛𝐺𝑑−𝐶𝑜 ̅̅̅̅̅̅̅̅̅|𝑑𝐺𝑑−𝐶𝑜|\n𝑟𝐺𝑑−𝐶𝑜 ̅̅̅̅̅̅̅̅2)](1) \nWhere 𝑛̅ is the average number of nearest neighbor s around all atoms, 𝑛𝐴−𝐵̅̅̅̅̅̅̅ is the average \nnumber of atom s A that are nearest neighbor s to atom B, 𝑟𝐴−𝐵̅̅̅̅̅̅ is average distance between \natom s A and nearest neighbor ing atom B. The 2\nπ factor comes from averaging of the cross product \n𝒔𝒊×𝒔𝒋 in DMI energy. \nFor 5 nm GdCo , with d Gd-Co < 0, dCo-Co < 0.25 x 10-22 J, only ferrimagnetic states are observed. At dCo-\nCo > 1.0 x 10-22 J, skyrmions are elongated due to boundary effect in the simulation or stripes states \nare observed. The range of DMI, where skyrmions are found, is smaller compared to calculation \nby Cort et al.21. This is due to a reduction in anisotropy an d exchange stiffness in GdCo. With less \nDMI energy required to create skyrmions, smaller DMI value is needed to create skyrmions and \nstripes in FiM. Furthermore, experiment results have measured DMI value greater than 1 mJ/m2 \nonly at ordered FM/heavy metal interface50-56. The DMI value at amorphous FiM/heavy metal \nremains unknown . Due to disorder nature of amorphous materials, the DMI value in amorphous \nFiM can be much smaller than the DMI value observed in ordered FM. \nAs shown in Fig. 4, with ferromagnetic DMI (d Gd-Gd and d Co-Co) that are positive, two scenarios of \nAFM DMI (d Gd-Co) are considered. At 300 K, in all thicknesses, larger DMI is needed to form \nskyrmions with positive dGd-Co than with negative dGd-Co. In 5 nm sample, D = 0.5 5 mJ/m2 is needed \nto stabilize skyrmions with dGd-Co > 0. In comparison, with dGd-Co < 0, a smaller DMI of D = 0.21 \nmJ/m2 is needed to stabilize skyrmions. Similar behaviors are also found in 10 nm and 15 nm \nsamples. With dGd-Co > 0, the smallest skyrmions are found at D = 1.26 m J/m2 in 10 nm sample and \nD = 2.31 m J/m2 in 15 nm sample. On the other hand, with dGd-Co < 0, the smallest skyrmions are \nfound at D = 0.84 m J/m2 in 10 nm sample and D = 1.68 m J/m2 in 15 nm sample. \nTo understand such intriguing behavior in a FiM, the in-plane spin configurations and the chirality \nof the skyrmion wall are investigated. Fig. 5 summarizes the chirality of the skyrmion wall s in the \nCo sublattice . Using dGd-Gd, dCo-Co > 0 and dGd-Co < 0, in the Co sublattice , the spins are turning in \ncounter -clockwise direction across the skyrmion wall . For Gd sublattice, the spins in the skyrmion \nwall are also turning counter -clockwise. This can be explained by the DMI in the system . AFM \ncoupling s between Gd and Co align the spins of Gd and Co in nearly antiparallel direction s, except \na small canting due to the presence of DMI. With dGd-Gd and dCo-Co > 0, turning counter -clockwise 4 \n is energetically favorable . However , with dGd-Gd, dCo-Co > 0 and dGd-Co > 0, the chirality of the \nsimulated skyrmion wall is found to be opposite. The DMI torque between the AFM pairs now \noppose s the DMI torques with in each sublattice . In the presence of a strong er inter -sublattice \nDMI torque , the spins in each sublattice now turn clockwise across the skyrmion wall. \nTo better illustrate the change in ch irality, the total DMI energies between Co -Co, Gd -Gd and Gd -\nCo are computed using the equilibrium configurations at 0 K. Table 1 summarizes the sign of the \ntotal DMI energies for different nearest neighbor pair s. With dGd-Gd, dCo-Co > 0 and dGd-Co> 0, spins \nare turning counter -clockwise . With this configuration, the total DMI energy between Gd -Gd pair \nEDMI(Gd-Gd) and Co-Co pair E DMI(Co-Co) are negative, and the total DMI energy between Gd and \nCo pair E DMI(Gd-Co) is also negative . This means that with dGd-Gd, dCo-Co > 0, it is energetically \nfavorable for spins to turn counterclockwise. On the other hand, with dGd-Gd, dCo-Co > 0 and dGd-Co > \n0, spins are revea led to turn clockwise from the simula ted configurations. As a result of the sign \nchange in chirality, EDMI(Gd-Gd) and E DMI(Co-Co) become positive . On the other hand, EDMI(Gd-Co) \nremains negative , bec ause both chirality and d Gd-Co change s sign. This implies that it is \nenergetically favorable for Gd -Co pair to turn clockwise across, but it is energetically unfavorable \nfor Gd -Gd and Co-Co pairs to do so. In other word , AFM DMI d Gd-Co is able to overcome \nferromagnetic DMI d Gd-Gd and dCo-Co, resulting in energy favorable configuration s for Gd -Co pairs. \nTo summarize , in a FiM, if the DMI of ferromagnetic pair and AFM pair have the same sign, a \ncancellation of DMI occurs because it is preferable for a ferromagnetic pair to turn in the opposite \ndirection of an AFM pair. No cancellation occurs if the DMI of ferromagnetic pair and AFM pair \nhave the opposite sign. These also explain the differences in size of skyrmion between d Gd-Co < 0 \nand d Gd-Co > 0. The dGd-Co < 0 scenario has larger skyrmions because both ferromagnetic and AFM \npairs are contributing to the formation of a skyrmion, which means the DMI effect is stronger \noverall. \nTo investigate the minimal size of room temperature skyrmions in GdCo, D -K phase diagrams with \nexponential ly decaying DMI at 300 K are simulated for 5, 10 and 15 nm GdCo films. In this section, \nwe focus on the d Gd-Co < 0 scenario . Since energy barrier is a function of exchange stiffness and \nthickness18, the minimal skyrmions size found in d Gd-Co < 0 scenario can also apply to d Gd-Co > 0 \nscenario, except a larger DMI is required . For each thi ckness, anisotropy ranges from 0.05 x 105 \nJ/m3 to 4 x 105 J/m3 are investigated. Experimentally, GdCo h as anisotropy in the order of 104 \nJ/m3.25,36 For DMI, larger interfacial DMI is explored in thicker samples, because a s thickness \nincreases, the average DMI decreases, and larger interfacial DMI is needed to stabilize skyrmions. \nIn 5 nm samples, interfacial DMI of 0 to 2 mJ/m2, which corresponds to dCo-Co of 0 to 2.38 x 1022 J, \nare investigated . Fig. 6 (a) shows the D -K phase diagram of 5 nm GdCo at 300 K. In 5 nm GdCo, \nskyrmions range from 12 nm to 40 nm are stabilized in the simulated range of interfacial DMI and \nanisotropy. Lines of 15 to 30 nm indicate the size of skyrmions at various DMI and anisotropy. As \nDMI decreases or anisotropy increase s, skyrmions become smaller and eventually collapse into \nFiM states. At the opposite side of D -K diagram, with large DMI and small anisotropy , skyrmions \nlarger than 40 nm becomes elongated or collapsed due to the boundary of the simula tion space \n(50.7 nm x 50.7 nm). This elongation of skyrmions was also seen earlier in Fig. 4 at large DMI \nvalues. Overall, for a given anisotropy, as interfacial D MI increases from 0 to 2.0 mJ/m2, the \nequilibrium configuration goes from FiM to skyrmions, then to stripes. For a fixed DMI, as \nanisotropy increases , size of skyrmions decreases, and finally, skyrmions collapse into FiM states. \nThese behavior of skyrmion s in FiM GdCo as a function of DMI and anisotropy is the same as what \nhas been observed in a ferromagnet17,18. 5 \n For 10 nm and 15 nm GdCo, DMI of 0 to 3 mJ/m2 (dCo-Co of 0 to 3.57 x 1022 J) and 0 to 4 mJ/m2 (dCo-\nCo of 0 to 4.76 x 1022 J) are explored respectively. The overall trend of skyrmions as a function of \nDMI and anisotropy in 10 nm and 15 nm GdCo are identical to that of 5 nm GdCo , where increase \nin DMI leads to larger skyrmions, and increase in anisotropy results in smaller skyrmions. However, \none difference in thicker sample s from 5 nm sample is that ultrasmall skyrmions as small as 7 nm \nare stable in room temperature. For both 10 nm and 15 nm GdCo, there is a region of DMI and \nanisotropy where ultrasmall skyrmions are stablized . In 10 nm GdCo, ultrasmall skyrmions are \nfound in the region of DMI ranges from 0.8 to 1.0 mJ/m2 and anisotropy ranges from (0.1 to 0.8) \nx 105 J/m3. For 15 nm GdCo, this region lays within DMI ranges from 1.5 to 1.8 mJ/m2 and \nanisotropy ranges from (0.1 to 1.0) x 105 J/m3. For both 10 nm and 15 nm GdCo, the anisotropy \nfalls within the same range as what has being measured experimentally in GdCo25,36, which is in \nthe order of 104 J/m3. However, the interfacial DMI is less than what has been observed at a Pt \ninterface. Ab-inito calculation has found Interfacial DMI of up to 12 mJ/m2 is reported at a n ideal \nPt/Co interface30. On the other hand, the interfacial DMI measured in Co/Pt and other Co -alloy/Pt \nfilms are around 1.2 to 1.5 mJ/m2. 52-54 Thus, some reduction s of DMI from that of Pt are needed \nto experimentally obtain ultrasmall skyrmion in 5 and 10 nm GdCo films. Reduction of DMI can be \nobtained by sandwiching GdCo between two Pt layers with one Pt layer being diluted by other \nelements. Since GdCo is amorphous, we have more flexibility of tuning the underlayer and the \ncapping layer of a multilayer sandwich. With its intrinsic anisotropy and flexibility , GdCo films are \npromising materials to obtain ultrasmall skyrmions at room temperature through DMI tuning . \nFor device applications, especially in thicker films, we will also need to consider the growth of \nskyrmions away from the interface. With decaying DMI away from the interface, spins at the top \nof a thicker sample experience effectively zero DMI. Without DM I, one might expect spins near \nthe top to align parallel for FM neighbors and antiparallel for AFM neighbors, and skyrmions to \ndisappear far away from the interface. If skyrmions collapse far away from the interface, the \nreliability of such memory devices would be vastly reduced. To investigate whether skyrmions \nremain robust in thicker samples, a numerical tomography is employed to image simulated \nultrasmall skyrmions at 300 K. Fig. 7 shows the numerical tomography plot of a ultrasmall \nskyrmion in 10 nm Gd Co. This skyrmion corresponds to D = 0.84 mJ/m2 and K = 0.3 x 105 J/m3. The \nsame skyrmion was shown in Fig. 4(b) and as the smallest skyrmions (Star Symbol) at K = 0.3 x 105 \nJ/m3 in Fig. 6(b). In the 3D plots at the center of Fig. 7, color s are made to be somewhat \ntransparent to reveal the skyrmions structure near the center. For Co sublattice, red to orange \ncolor shows that most of the spins are pointing down . A region of green and blue that appears \nnear the center corresponds to the simula ted skyrmion at 300 K. As evidenced by the columnar \ndistribution of blue color, the skyrmion retain s a uniform columnar growth from the bottom to \nthe top. Columnar distribution of skyrmion is also found in Gd sublattice , where a column of red \nis distribute d uniformly from the bottom to the top . This feature can be understood in terms of \nthe large magnetic exchange length >20 nm due to the low magnetization in the ferrimagnet . \nTo further demonstrate the uniform columnar distribution of skyrmion, in -plane and out -of-plane \ncross sections of the skyrmion are also plotted in Fig. 7. On the left of Fig. 7, in-plane cross section \nof spin configuration within 0.5 nm of the interface and 0.5 nm of the top are mapped. The \nskyrmion s at the interface and near the top have identical size an d shape. Compare to the \nmapping of spin configuration s in Fig. 4(b), size of the skyrmion remain the same. This shows that \nthe size of skyrmions rema in the uniform throughout a sample. On the right side of Fig. 7, out -of-\nplane cross section s are shown for Gd and Co sublattice s. The blue color in Co sublattice and the \nred color in Gd sublattice correspond to the center of the skyrmion. For both sublattice, out -of-\nplane cross sections show a columnar distribution of skyrmion from the bottom in terface to the 6 \n top. These results provide important evidences that skyrmion remain robust through a thicker \nsample, and further support of using thicker GdCo samples to increase skyrmion stability at room \ntemperature . \nConclusions \nUsing atomistic stochastic LLG simulations, ultrasmall skyrmions are shown to be stable at room \ntemperature in ferri magnetic GdCo. Despite the rapid decay of Dzyaloshinskii Moriya interaction \n(DMI) away from the interface, a rea listic range of DMI values is seen to stabilize skyrmions in \nGdCo films as thick as 15 nm irrespective of the sign of DMI between antiferromagnetic coupled \nGd and Co, Furthermore, the low DMI values needed to form ultrasmall skyrmion in GdCo indicate \nopportunity for design ing magnet ic materials to host ultrasmall Neel skyrmions . Through \ntomography of an ultrasmall skyrmion in 10 -nm thick GdCo film, it is discovered that the skyrmion \nassumes a columnar configuration that extends uniformly across the film thickness despite having \nnear zero DMI far away from the interface. These findings argue for using thicker magnetic films \nto host ultrasmall skyrmions, providing an important strategy for developing high density and high \nefficiency skyrmion based devices. \nMethods \nThe classical atom istic Hamiltonian H in Eq. (1) is employed to investigate magnetic textures in \namorphous FiMs. \n𝐻=−1\n2∑ 𝐽𝑖𝑗𝒔𝒊∙𝒔𝒋\n<𝑖,𝑗>−1\n2∑ 𝐷𝑖𝑗∙(𝒔𝒊×𝒔𝒋)\n<𝑖,𝑗>−𝐾𝑖(𝒔𝒊∙𝑲𝒊̂)2 \n−𝜇0𝜇𝑖𝑯𝒆𝒙𝒕∙𝒔𝒊−𝜇0𝜇𝑖𝑯𝒅𝒆𝒎𝒂𝒈 ∙𝒔𝒊 (2) \nwhere 𝒔𝒊,𝒔𝒋 are the normalized spins and 𝜇𝑖,𝜇𝑗 are the atomic moments at sites i, j respectively. \nThe atomic moment is absorbed into the exchange constant, 𝐽𝑖𝑗=𝜇𝑖𝜇𝑗𝑗𝑖𝑗, the DMI interaction \n𝑫𝒊𝒋=𝜇𝑖𝜇𝑗𝒅𝒊𝒋, which is proportional to ri x rj, the po sitional vector between the atoms i, j and the \ninterface , and the effective anisotropy 𝐾𝑖=𝜇𝑖𝑘𝑖. 𝑯𝒆𝒙𝒕 and 𝑯𝒅𝒆𝒎𝒂𝒈 are the external field and \ndemagnetization field respectively. \nOnly nearest neighbor interactions are considered in the exchange and DMI interactions. Periodic \nboundary conditions are enforced in the x and y direction s. \nTo find the ground state, the spins are evolved under the following stochastic Landau -Lifshitz -\nGilbert (LLG) equation, \n𝑑𝑴\n𝑑𝑡=−𝛾\n1+𝛼2𝑴×(𝑯𝒆𝒇𝒇+𝝃)−𝛾𝛼\n(1+𝛼2)𝑀𝑠𝑴×[𝑴×(𝑯𝒆𝒇𝒇+𝝃)] (3) \nwhere 𝛾 is the gyromagnetic ratio, 𝛼 is the Gilbert damping constant, 𝑯𝒆𝒇𝒇 is the effective field, \n𝝃 is the Gaussian white noise term for thermal fluctuations and 𝑀𝑠 is the saturation magnetization. \nThe parameters used in our simulation are listed in Table 2. Exchange couplings 𝐽𝑖𝑗are calibrated \nbased on Ostler et al.49 to maintain the same Curie temperature and compensation temperature \nfor a given compensation. At 300 K, the magnetization of Gd 25Co75 is 5 x 104 A/m. Anisotropy \nenergy is determined based on Hansen et al.36 7 \n To incorporate the amorphous short range order, an amorphous structure of a 1.6 nm x 1.6 nm x \n1.6 nm box containing 250 atoms is generated from ab initio molecular dynamics calculations by \nSheng et al .58. The composition used in the simulation is Gd 25Co75. Fig. 1 shows a plot of RE and \nTM atoms in the amorphous structure. For a 4.8 nm thick sample, replicas of th is box (32 x 32 x 3) \nare placed next to each other to expand the simulated sample to 50.7 nm x 50.7 nm x 4.8 nm and \n768000 atoms. On average, we find that each Co atom has 6.8 Co neighbor s and 4.1 Gd neighbor s, \nwhile each Gd atom has 11.7 Co neighbors and 3.5 Gd neighbor s. We have also employed a FC C \nlattice to study skyrmions in GdCo. We found that with the same compensation temperature and \nmagnetization, a larger DMI is needed to stabilize skyrmion in a FCC lattice structures than \namorphous structure. This is because the overall effectiveness of DM I is affected by the structure. \nOnly results using the amorphous structure are shown herein. \nIn the simulations, the initial states are skyrmion of 20 nm based on the 2 -pi model18. Various \ninitial states, includes random initial states and 10 -30 nm skyrmions, have been tested and found \nto produce the same final states. The size of skyrmions are defined as the diameter for which M z \n= 0. Since skyrmions are not perfectly symmetric, size of skyrmion is the avera ge diameter . \nData Availability \nThe datase ts generated during and/or analyzed during the current study are available from the \ncorresponding author on reasonable request. \nReferences: \n1. Rößler, U. K., Bogdanov, A. N. & Pfleiderer, C. Spontaneous skyrmion ground states in \nmagnetic metals. Nature 442, 797–801 (2006). \n2. Yu, X. Z. et al . Real -space observation of a two -dimensional skyrmion crystal. Nature 465, \n901–904 (2010). \n3. Yu, X.Z. et al . Skyrmion flow near room temperature in an ultralow current density. Nat. \nCommun. 3, 988 (2012). \n4. Nagaosa, N. & Tokura , Y. Topological properties and dynamics of magnetic skyrmions . Nat. \nNanotechnol. 8, 899 -911 (2013). \n5. Sampaio, J., Cros, V., Rohart, S., Thiaville A. & Fert, A. Nucleation, stability and current -\ninduced motion of isolated magnetic skyrmions in nanostructur es. Nat. Nanotech nol. 8, 839 -\n844 (2013). \n6. Jiang, W. et al. Blowing magnetic skyrmion bubbles. Science 349, 283 –286 (2015). \n7. Büttner, F. et al. Dynamics and inertia of skyrmionic spin structures. Nat. Phys. 11, 225 -228 \n(2015). \n8. Romming, N. et al. Writing and deleting single magnetic skyrmions. Science 341, 636 –639 \n(2013). \n9. Romming, N., Kubetzka, A., Hanneken, C., von Bergmann, K. & Wiesendanger, R. Field -\ndependent size and shape of single magnetic skyrmions. Phys. Rev. Lett. 114, 177203 (2015). \n10. Boulle, O. et al. Room -temperature chiral magnetic skyrmions in ultrathin magnetic \nnanostructures. Nat. Nanotech nol. 11, 449 -454 (201 6). \n11. Zhang, X., Ezawa, M. & Zhou Y. Magnetic skyrmion logic gates: conversion, duplication and \nmerging of skyrmions. Sci. Rep. 5, 9400 (2015). \n12. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A. , Neubauer, A., Georgii, R. & Böni, \nP. Skyrmion Lattice in a Chiral Magnet . Science 323, 915 -919 (2009). \n13. Yu, X.Z. et al . Near room -temperature formation of a skyrmion crystal in thin -films of the 8 \n helimagnet FeGe . Nat. Mater. 10, 106 –109 (2011). \n14. Tolley, R., Montoya, S.A. & Fullerton, E.E. Room -temperature observation and current \ncontrol of skyrmions in Pt/Co/Os/Pt thin films . Phys. Rev. Mater. 2, 044404 (2018). \n15. Woo, S. et al. Observation of room -temperature magnetic skyrmions and their current -\ndriven dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501 –506 (2016). \n16. Soumyanarayanan, A. et al. Tunable room -temperature magnetic skyrmions in Ir/Fe/Co/Pt \nmultilayers. Nat. Mater. 16, 898 –904 (2017). \n17. Siemens, A., Zhang, Y., Hagemeister, J., Vedmedenko, E.Y. & Wiesendanger, R. Minimal \nradius of magnetic skyrmions: statics and dynamics. New. J. Phys. 18, 045021 (2016). \n18. Büttner, F., Lemesh I. & Beach G.S.D. Theory of isolated magnetic skyrmions: From \nfundamentals to room temperature applications. Sci. Rep. 8, 4464 (2018). \n19. Jiang, W. et al. Direct observation of the skyrmion Hall effect . Nat. Phys. 13, 162 -169 (2017). \n20. Litzius, K. et al. Skyrmion Hall effect revealed by direct time -resolved X -ray microscopy . Nat. \nPhys. 13, 170 -175 (2017). \n21. Fert, A., Cros, V., Sampaio, J. Skyrmions on the track. Nat. Nanotechnol. 8, 152 -156 (2013). \n22. Tomasello, R. et al . A strategy for the design of skyrmion racetrack memories. Sci. Rep. 4, \n6784 (2014). \n23. Woo, S. et al. Current -driven dynamics and inhibition of the skyrmion Hall effect of \nferrimagnetic skyrmions in GdFeCo films. Nat. Commun. 9, 959 (2018). \n24. Lee, J. C. T. et al. Synthesizing skyrmion bound pairs in Fe -Gd thin films. Appl. Phys. Lett. 109, \n(2016). \n25. Caretta, L. et al. Fast current -driven domain walls and small skyrmions in a compensated \nferrimagnet . Nat. Nanotechnol. 13, 1154 -1160 (2018). \n26. Dzyaloshinsky, I. A thermodynamic theory of weak ferromagnetism of antiferromagnetics . J. \nPhys. Chem. Solids 4, 241 –255 (1958). \n27. Moriya, T. Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120, \n91–98 (1960). \n28. Belmeguenai, M. et al. A. Interfacial Dzyalo shinskii –Moriya interaction in perpendicularly \nmagnetized Pt/Co/AlOx ultrathin films measured by Brillouin light spectroscopy. Phys. Rev. B \n91, 180405(R) (2015). \n29. Nembach, H.T., Shaw, J.M, Weiler, M., Jué E. & Silva, T.J. Linear relation between Heisenberg \nexchange and interfacial Dzyaloshinskii –Moriya interaction in metal films . Nature Phys. 11, \n825-829 (2015). \n30. Yang, H., Thiaville, A., Rohart, S., Fert, A. & Chshiev, M. Anatomy of Dzyaloshinskii -Moriya \nInteraction at Co/Pt Interfaces . Phys. Rev. Lett. 118, 219901 (2017) \n31. Belmeguenai, M. et al. Thickness Dependence of the Dzyaloshinskii -Moriya Interaction in \nCo2FeAl Ultrathin Films: Effects of Annealing Temperature and Heavy -Metal Material . Phys. \nRev. Appl. 9, 044044 (2018). \n32. Dirks, A. G. & Leamy, H. J. Colum nar microstructure in vapor -deposited thin films. Thin Solid \nFilms 47, 219–233 (1977). \n33. Leamy, H. J. & Dirks, A. G. Microstructure and magnetism in amorphous rare -earth -\ntransition -metal thin films. II. Magnetic anisotropy. J. Appl. Phys. 49, 3430 (1978). \n34. Harris, V. G., Aylesworth, K. D., Das, B. N., Elam, W. T. & Koon, N. C. Structural origins of \nmagnetic anisotropy in sputtered amorphous Tb -Fe films. Phys. Rev. Lett. 69, 1939 –1942 \n(1992). \n35. Harris, V. G. & Pokhil, T. Selective -Resputtering -Induced Perpendicul ar Magnetic Anisotropy 9 \n in Amorphous TbFe Films. Phys. Rev. Lett. 87, 67207 (2001). \n36. Hansen, P., Clausen, C., Much, G., Rosenkranz, M. & Witter, K. Magnetic and magneto -\noptical properties of rare -earth transition -metal alloys containing Gd, Tb, Fe, Co. J. Appl. \nPhys. 66, 756–767 (1989). \n37. Kim, K -J. et al. Fast domain wall motion in the vicinity of the angular momentum \ncompensation temperature of ferrimagnets . Nature Materials 16, 1187 -1192 (2017). \n38. Stanciu, C. D. et al. All-optical magnetic recording with circularly polarized light. Phys. Rev. \nLett. 99, 47601 (2007). \n39. Savoini, M. et al. Highly efficient all -optical switching of magnetization in GdFeCo \nmicrostructures by interference -enhanced absorption of light. Phys. Rev . B 86, 140404(R) \n(2012). \n40. Ostler, T.A. et al. Ultrafast heating as a sufficient stimulus for magnetization reversal in a \nferrimagnet . Nat. Commun. 3, 666 (2012). \n41. Hassdenteufel, A. et al. Thermally assisted all -optical helicity dependent magnetic switching \nin amorphous Fe 100-xTbx alloy films. Adv. Mater. 25, 3122 –3128 (2013). \n42. Kirilyuk, A., Kimel, A. V. & Rasing, T. Ultrafast optical manipulation of magnetic order. Rev. \nMod. Phys. 82, 2731 –2784 (2010). \n43. Kirilyuk, A., Kimel, A. V. & Rasing, T. Laser -induced magnetization dynamics and reversal in \nferrimagnetic alloys. Rep. Prog. Phys. 76, 026501 (2013) \n44. Kimel, A. V. All -optical switching: Three rules of design. Nat. Mater. 13, 225–226 (2014). \n45. Magnin, S. et al. Engineered materials for all -optical helicity -dependent magnetic switching. \nNat. Mater. 13, 286-292 (2014). \n46. Ostler, T. A. et al. Crystallographically amorphous ferrimagnetic alloys: Comparing a \nlocalized atomistic spin model with experiments. Phys. Rev. B 84, 24407 (2011). \n47. Radu, I. et al. Transient ferromagnetic -like state mediating ultrafast reversal of \nantiferromagnetically coupled spins. Nature 472, 205–208 (2011). \n48. Ellis, M. O. A., Ostler, T. A. & Chantrell, R. W. Classical spin model of the relaxatio n dynamics \nof rare -earth doped permalloy. Phys. Rev. B 86, 174418 (2012). \n49. Evans, R. F. L. et al. Atomistic spin model simulations of magnetic nanomaterials. J. Phys. \nCondens. Matter 26, 103202 (2014). \n50. Atxitia, U., Nieves, P. & Chubykalo -Fesenko O. Landau -Lifshitz -Bloch equation for \nferrimagnetic materials. Phys. Rev. B 86, 104414 (2012). \n51. Chen, G. et al . Tailoring the chirality of magnetic domain walls by interface engineering. Nat. \nCommun. 4, 2671 (201 3). \n52. Hrabec, A. et al . Measuring and tailoring the Dzyaloshinskii -Moriya interaction in \nperpendicularly magnetized thin films. Phys. Rev. B 90, 020402(R) (2014). \n53. Stashkevich, A.A. et al. Experimental study of spin -wave dispersion in Py/Pt film structures in \nthe presence of an interface Dzyaloshinskii -Moriya interaction . Phys. Rev. B 91, 214409 (2015). \n54. Ma, X. et al. Interfacial control of Dzyaloshinskii -Moriya interaction in heavy \nmetal/ferromagnetic metal thin film heterostructures . Phys. Rev. B 94, 180408( R) (2016). \n55. Tacchi, S. et al . Interfacial Dzyaloshinskii -Moriya Interaction in Pt/CoFeB Films: Effect of the \nHeavy -Metal Thickness . Phys. Rev. Lett. 118, 147201 (2017). \n56. Cho, J. et al. The sign of the interfacial Dzyaloshinskii –Moriyainteraction in ultrathin \namorphous and polycrystalline magnetic films . J. Phys. D: Appl. Phys. 50, 425004 (2017). \n57. Simon, E., Rózsa, L., Palotás, K. & Szunyogh , L. Magnetism of a Co monolayer on Pt(111) \ncapped by overlayers of 5d elements: A spin -model study . Phys. Rev. B 97, 134405 (2018). 10 \n 58. Sheng, H.W., Luo, W.K., Alamgir, F.M., Bai, J.M., & Ma, E. Atomic packing and short -to-\nmedium -range order in metallic glasses. Nature 439, 419 -425 (2006). \n59. Zhang, X., Zhou, Y., & Ezawa, M. Antiferromagnetic Skyrmion: Stability, Creation and \nManipulation . Sci. Rep. 6, 24795 (2016). \n60. Baker, J., & Tretiakov, O. A., Static and Dynamical Properties of Antiferromagnetic Skyrmions \nin the Presence of Applied Current and Temperature . Phys. Rev. Lett. 116, 147203 (2016). \n61. Kim, K -J. et al. Fast domain wall motion in the vicinity of the angular momentum \ncompensation temperature of ferrimagnets . Nature Materials 16, 1187 -1192 (2017). \nAcknowledgements: \nThis work was supported by the DARPA Topological Excitations in Electronics (TEE) program (grant \nD18AP00009). The content of the information does not necessarily reflect the position or the \npolicy of the Government, and no official endorsement should be inferred. Approved for public \nrelease; distribution is unlimited . This work was partially supporte d by NSF -SHF-1514219. \nAdditional Information \nAuthor Contributions \nS.J.P. conceived the project and supervised the simulation, C.T.M. performed the simulation, \nY.X. assisted in the simulation and provided comments, A .W.G. discussed skyrmion stability and \nsuggested improve ment to the manuscript , H.S. provided the amorphous structure from ab \ninitio molecular dynamics calculations. \nCompeting Interests \nThe authors declare no competing interests. \n \n \n \n \n 11 \n \nFigure 1 Amorphous st ructure of RE 25TM 75 from ab initio molecular dynamics calculations . Red \natoms are rare -earth, and blue atoms are transition -metal. \n \nFigure 2 Simulated saturation magnetization vs. temperature of amorphous Gd 25Co75. The \ncompensation temperature of amorphous Gd 25Co75 is near 250 K , and the magnetization is small \nat room temperature. \n12 \n \nFigure 3 Exponential decay DMI in 5 nm GdCo as function of distance from bottom interface (z) . \nIn this model, DMI remains constant (D0) within 0.35 nm of the bottom int erface, as indicated by \nthe red line. Away from the interface, the strength of DMI decays exponentially as shown. \n13 \n \nFigure 4 Color mapping of equilibrium spin configurations for various DMI values (exponentia lly \ndecay ing DMI) at 300K for dGd-Co > 0 and dGd-Co < 0 in (a ) 5 nm, (b) 10 nm, and (c) 15 nm GdCo . \nOut-of-plane component s of reduced magnetizations (mz) are mapped in the x -y plane using the \ncolor bar shown in (c). Ultrasmall skyrmions are revealed in 10 nm and 15 nm GdCo samples. \n14 \n \nFigure 5 Simulated s kyrmion configurations of Co sublattice for dGd-Co < 0 and dGd-Co > 0 with metal \ninterface at the bottom . (Top) A n overhead view of simulated skyrmion configurations for dGd-\nCo < 0 and dGd-Co > 0. (Bottom) For dGd-Co < 0, the skyrmion wall is turning counter -clockwise. The \nd(FM) vector is pointing in the opposite direction of Si x Sj. EDMI (FM)= dij · (Si x Sj ) is negative, \nwhich is favorable. For dGd-Co > 0, the skyr mion wall is turning clockwise. The d(FM) vector and Si x \nSj are pointing in the same direction, resulting in positive EDMI (FM) . Identical signs of the DMI \nenergy are also found in the Gd sublattice. \nScenario EDMI(Gd-Gd) EDMI(Co-Co) EDMI(Gd-Co) \ndGd-Gd, dCo-Co > 0, dGd-Co < 0 - - - \ndGd-Gd, dCo-Co > 0, dGd-Co > 0 + + - \nTable. 1 Sign of total DMI energy E DMI computed from equilibrium spin configurations at 0 K. \n15 \n \nFigure 6 D-K phase diagram of (a) 5 nm, (b) 10 nm and (c) 15 nm GdCo at 300 K with dGd-Co < 0. \nStar corresponds to smallest skyrmions simulated at K = 0 .3 x 105 J/m3. Ultrasmall skyrmions are \nrevealed in 10 nm and 15 nm GdCo. Due to limits of simulation space (50.7 nm x 50.7 nm), with \nperiodic boundary conditions in x -y direction, large skyrmions (> 40 nm) become either elongated \nor collapsed. \n \n16 \n \n \nFigure 7 Tomograph of a simulated ultrasmall skyrmion in 10 nm GdCo at 300 K . It reveals \ncolumnar skyrmion distribution throughout the 10 nm GdCo sample. The figure shows Co -\nsublattice spins (top box), Gd -sublattice spins (bottom box), in -plane cross sections of near the \ntop and bottom interface (left), and out -of-plane cross sections (right). \nParameter Value \nGyromagnetic ratio (ϒ) 2.0023193 \nGilbert Damping (α) 0.05 \nGd moment ( μGd) 7.63 μB \nCo moment ( μCo) 1.72 μB \nGd-Gd exchange constant (J Gd-Gd) 1.26 x 10-21 J \nCo-Co exchange constant (J Co-Co) 3.82 x 10-21 J \nGd-Co exchange constant (J Gd-Co) -1.09 x 10-21 J \nMagnetic Field (H) 0.01 T \nTable. 2 Values of parameters used in the simulation. \n \n" }, { "title": "2309.01965v1.Strong_and_nearly_100_____spin_polarized_second_harmonic_generation_from_ferrimagnet_Mn___2__RuGa.pdf", "content": "arXiv:2309.01965v1 [cond-mat.mtrl-sci] 5 Sep 2023Strong and nearly 100 %spin-polarized second-harmonic generation from ferrimag net\nMn2RuGa\nY. Q. Liu,1M. S. Si∗,1and G. P. Zhang†2\n1School of Materials and Energy, Lanzhou University, Lanzho u 730000, China\n2Department of Physics, Indiana State University, Terre Hau te, IN 47809, USA\n(Dated: September 6, 2023)\nSecond-harmonic generation (SHG) has emerged as a promisin g tool for detecting electronic and\nmagnetic structures in noncentrosymmetric materials, but 100% spin-polarized SHG has not been\nreported. In this work, we demonstrate nearly 100% spin-pol arized SHG from half-metallic ferri-\nmagnet Mn 2RuGa. A band gap in the spin-down channel suppresses SHG, so t he spin-up channel\ncontributes nearly all the signal, as large as 3614 pm/V abou t 10 times larger than that of GaAs.\nIn the spin-up channel, χ(2)\nxyzis dominated by the large intraband current in three highly d ispersed\nbands near the Fermi level. With the spin-orbit coupling (SO C), the reduced magnetic point group\nallows additional SHG components, where the interband cont ribution is enhanced. Our finding is\nimportant as it predicts a large and complete spin-polarize d SHG in a all-optical spin switching\nferrimagnet. This opens the door for future applications.\nPACS numbers:\nI. INTRODUCTION\nThe interaction between an intense optical field and a\nmaterial is always fascinating. This gave birth to non-\nlinear optics [1, 2]. Second-harmonic generation (SHG),\na special case of sum frequency generation, has received\nenormous attention worldwide. SHG only exists in non-\ncentrosymmetricmaterialswith brokeninversionsymme-\ntryI[3], while it is absent in centrosymmetric systems.\nIn general, impurities and surfaces introduced in a mate-\nrial can break I. For instance, Iof the NV center is bro-\nken by introducing nitrogen-vacancies in diamond [4–7].\nMoreover, a few layers of crystal, created by mechanical\nexfoliation, exhibit different symmetry properties. Odd\nlayers of MoS 2and h-BN belong to the noncentrosym-\nmetric space group, different from their bulk, can also\ngenerate SHG [8–10]. So far, most of the materials stud-\nied are nonmagnetic. For magnetic materials, magnetic\norder can break time reversal symmetry T. The sizable\nSHG appears in the antiferromagnetic (AFM) CrI 3and\nthe even septuple layers of MnBi 2Te4[11–13], where the\nAFM orderingbreaks I. In these two cases, SOC destroys\nthe symmetry of band structure thereby enhancing SHG.\nAlthough nearly 100% spin polarization at the Fermi\nlevel is observed in materials such as half-metal Cr 2O3\n[14], they possess I, where only the odd-order harmon-\nics are observed. Until now, little is known about 100%\nspin-polarized SHG in a half-metallic ferrimagnet.\nIn this work, we predict a strong SHG signal from the\nhalf-metallic Heusler Mn 2RuGa. We show that a single\nspin channel mainly contributes to SHG in Mn 2RuGa.\nThe band gap in the spin-down channel is open and lim-\nits SHG, resulting in SHG mainly from the spin-up chan-\nnel. Surprisingly, χ(2)\nxyzreaches as large as 3614 pm/V for\nthe spin-upchannel, atleastanorderofmagnitudelarger\nthan that ofGaAs. It is found that the intrabandcurrent\ndominates this large χ(2)\nxyz, which originates from threehighly dispersive bands near the Fermi level. To con-\nfirm our conclusion, we remove the Ru atoms to obtain\nMn2Ga, where both the spin-up and spin-down channels\nare metallic. It is found that a highly dispersive band\nnear the Fermi level appears in the spin-down channel,\nwhich does enhance the SHG spectrum χ(2)\nxyz. With SOC,\nthe spin-up and spin-down channels arecoupled. As a re-\nsult, the restricted transitions in the spin-polarized case\nare now SOC-allowed between the flat valence and con-\nduction bands. The underlying physics stems from the\nreduced magnetic point group induced by SOC, where\nthe magnetization field is applied along the z-axis. This\ndirectly leads to the appearance of additional SHG com-\nponents such as χ(2)\nxxz, where the allowed interband tran-\nsitions play a role. Our study demonstrates that nearly\n100% spin-polarized SHG can detect the half-metallicity\nin Heusler alloy Mn 2RuGa.\nThe rest of the paper is arrangedas follows. In Sec. II,\nwe show our theoretical methods. Then, the results and\ndiscussions are given in Sec. III. Finally, we conclude our\nwork in Sec. IV.\nII. COMPUTATIONAL METHODS\nA. First-principle electronic structure calculations\nThe electronic structures of Mn 2RuGa are calculated\nwithin the first-principle density functional theory using\ntheprojector-augmentedwave(PAW)[15,16]method, as\nimplemented in the Vienna Ab intio Simulation Package\n(VASP) [17–20]. The generalized gradient approxima-\ntion (GGA) [21] is employed within the Perdew-Burke-\nErnzerhof (PBE) scheme as the exchange-correlation\nfunctional. We self-consistently solve the Kohn-Sham2\nequation\n/bracketleftbigg\n−¯h2\n2me∇2+Vne(/vector r)+e2\n4πǫ0/integraldisplayn(/vector r)\n|/vector r−/vector r′|d3/vector r′+Vxc(/vector r)/bracketrightbigg\n×ψn/vectork(/vector r) =εn/vectorkψn/vectork(/vector r).\n(1)\nThe first term is the kinetic energy and next three terms\nare the potential energy, the Coulomb, and the exchange\ninteractions, respectively. meis the electron mass and\nn(/vector r) is the electron density. ψn/vectork(/vector r) denotes the Bloch\nwave function of band nat crystal momentum /vectork, and\nεn/vectorkis the band energy. The cutoff energy is set to 500\neV. The structural optimizations and self-consistent are\ncarried out by Γ-centered k-point mesh of 15 ×15×15.\nThe density of states (DOS) is calculated using a denser\nk-point mesh of 21 ×21×21. In order to obtain accurate\nresults, we set the energy convergence less than 10−6eV.\nB. First-principle nonlinear optical response\ncalculations\nWe use the length gauge to compute SHG [22–\n24]. The nonlinear polarization is given by Pa(2ω) =\nχabc(2ω;ω,ω)Eb(ω)Ec(ω), whereχabcdenotes the SHG\nsusceptibility, and Eb(ω) is thebcomponent of the op-\ntical electric field at frequency ω. In general, χabccon-\ntains three major contributions: the interband transi-\ntionsχabc\ninter, the intraband transitions χabc\nintra, and the\nmodulation of interband terms by intraband terms χabc\nmod,\nand can be expressed as\nχabc(2ω;ω,ω) =χabc\ninter(ω)+χabc\nintra(ω)+χabc\nmod(ω),\n=χabc\n2ph,inter(ω)+χabc\n1ph,inter(ω)\n+χabc\n2ph,intra(ω)+χabc\n1ph,intra(ω)\n+χabc\nmod(ω),(2)\nwherethe subscripts2 phand1phrepresenttwo-andone-\nphoton transitions, respectively. The interband and in-\ntraband transitions are\nχabc\ninter(ω) =χabc\n2ph,inter(ω)+χabc\n1ph,inter(ω),\nχabc\nintra(ω) =χabc\n2ph,intra(ω)+χabc\n1ph,intra(ω).(3)\nThedetailedexpressions[25]ofthesefourtermsaregiven\nby\nχabc\n2ph,inter(ω) =e3\n¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlra\nnm(k)/braceleftbig\nrb\nml(k)rc\nln(k)/bracerightbig\nωln(k)−ωml(k)\n×2fnm\nωmn(k)−2ω−2iη,\n(4)\nχabc\n1ph,inter(ω) =e3\n¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlra\nnm(k)/braceleftbig\nrb\nml(k)rc\nln(k)/bracerightbig\nωln(k)−ωml(k)\n×/braceleftbiggfml\nωml(k)−ω−iη+fln\nωln(k)−ω−iη/bracerightbigg\n,\n(5)χabc\n2ph,intra(ω) =e3\n¯h2/integraldisplaydk\n4π3/summationdisplay\nnmra\nnm(k)/braceleftbig\n∆b\nmn(k)rc\nmn(k)/bracerightbig\nω2mn(k)\n×−8ifnm\nωmn(k)−2ω−2iη\n−e3\n¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlra\nnm(k)/braceleftbig\nrb\nml(k)rc\nln(k)/bracerightbig\nω2mn(k)\n×2fnm(ωln(k)−ωml(k))\nωmn(k)−2ω−2iη,\n(6)\nχabc\n1ph,intra(ω) =e3\n¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlra\nnm(k)/braceleftbig\nrb\nml(k)rc\nln(k)/bracerightbig\n×ωmn(k)/braceleftbiggfnl\nω2\nln(k)(ωln(k)−ω−iη)\n−flm\nω2\nml(k)(ωml(k)−ω−iη)/bracerightbigg\n.\n(7)\nWe know that χabc\nmodin Eq. (2) contributes little to SHG\nin comparison with χabc\ninterandχabc\nintra, and it has the form\nas\ni\n2ωχabc\nmod(2ω;ω,ω) =\nie3\n2¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlωnl(k)ra\nlm(k)/braceleftbig\nrb\nmn(k)rc\nnl(k)/bracerightbig\nfnm\nω2mn(k)(ωmn(k)−ω−iη)\n−ie3\n2¯h2/integraldisplaydk\n4π3/summationdisplay\nnmlωlm(k)ra\nnl(k)/braceleftbig\nrb\nlm(k)rc\nmn(k)/bracerightbig\nfnm\nω2mn(k)(ωmn(k)−ω−iη)\n+ie3\n2¯h2/integraldisplaydk\n4π3/summationdisplay\nnmfnm∆a\nnm(k)/braceleftbig\nrb\nmn(k)rc\nnm(k)/bracerightbig\nω2mn(k)(ωmn(k)−ω−iη).\n(8)\nIn Eqs. (4)-(8), ωmn(k) =ωm(k)−ωn(k), and the\nenergy of band nis ¯hωn. ∆a\nmn(k) =υa\nmm(k)−\nυa\nnn(k), whereυa\nnm(k) is theacomponent of the ve-\nlocity matrix elements, and fmn=f(¯hωm)−f(¯hωn),\nwheref(¯hωn) is the Fermi Dirac function. The ma-\ntrix elements of position operator rmn(k) is given by\nra\nmn(k) =−iυa\nnm(k)/ωmn(k), and/braceleftbig\nrb\nml(k)rc\nln(k)/bracerightbig\n=\n1\n2/bracketleftbig\nrb\nml(k)rc\nln(k)+rc\nml(k)rb\nln(k)/bracketrightbig\n. The matrix elements of\nposition operator rmn(k) are directly computed by the\nfirst-principle calculations, which accounts for the effect\nof non-local potentials [26]. The damping parameter η\nis set to 0.003 Hartree. In the realistic calculations, the\nnumber ofkpoints and energy bands affect the accuracy\nofχabc. A large number of kpoints are required to ob-\ntain an accurate NLO response, so a very dense k-point\nmesh of32 ×32×32is used. The number of energybands\nis set to 32 to converge the spin-polarized SHG spectra,\nand 64 for the SHG spectra with SOC.3\nIII. RESULTS AND DISCUSSIONS\nA. Crystal, electronic structures and\nspin-polarized SHG of Mn 2RuGa\n(a)\nHalf-metallic ferrimagnet0.0 0.5 1.0 1.5 2.0 2.5 3.030006000900012000\n0Spin-up Spin-dn(b)\n|\n\u0001xyz| (pm/V)(2)\n0 1 2 3 4 5 6100200300\n0(c)\nE (eV) GaAsMn2RuGa2 \u0002\n7 8zLaser\nDetector\nEF\nSpin-up Spin-dngap2\u0000 2\u0003\n\u0004\u0005\n\u0006400\n|\n\u0007xyz| (pm/V)(2)\nFIG. 1: (a) A sketch of the nearly 100% spin-polarized SHG\nin half-metallic ferrimagnets, where the contribution fro m the\nspin-up channel is significant, while that from the spin-dow n\nchannel is largely suppressed. (b) The absolute value of the\nSHG susceptibility χ(2)\nxyzfor Mn 2RuGa without SOC. The red\nsolid andblack dashed lines denote the spin-upand spin-dow n\nchannels, respectively. The damping parameter η= 0.003\nHartree is taken. (c) χ(2)\nxyzof GaAs is also given for compari-\nson.\nSHG appears in nonmagnetic materials with broken I,\nbut it does not distinguish spin. For ferrimagnetic ma-\nterials, both spin-up and spin-down channels contribute\nto SHG. However, SHG from two spin channels is differ-\nent, which is closely related to the spin-polarized band\nstructures near the Fermi level. For a half-metallic ferri-\nmagnet with nearly 100% spin polarization at the Fermi\nlevel, SHG mainly comes from one spin channel show-\ning metallicity, as schematically displayed in Fig. 1(a).\nMn2RuGa is such a half-metallic ferrimagnet with many\ndifferent structures [27–29]. We choose the most stable\nXAHeusler structure, which belongs to the space group\nF-43m [30], as shown in Appendix A. This structure is\nmore consistent with the experimental results [31]. The\nexperimental lattice parameters are a=b=c= 5.97\n˚A [32]. In Mn 2RuGa, there are two Mn atoms, Mn 1\nand Mn 2, with different Wyckoff positions 4a(0, 0, 0)\nand 4c(1/4, 1/4, 1/4). The positions of Ru and Ga are\n4d(3/4, 3/4, 3/4) and 4b(1/2, 1/2, 1/2), respectively.\nThe magnetic moments of two Mn atoms are M4a= 3.13\nµB, andM4c=−2.29µB, consistent with the prior\nstudy [30], and they are antiferromagnetically coupled.\nHere, the magnetic moment direction is along the zdi-\nrection. It is found that the total magnetic moment is\nMtot= 1.03µB, which satisfies the Slater-Paulingrule as\nreported in the literature [33]. This rule provides a sim-\nple relationshipbetween the totalmagneticmoment Mtot\nand the valence electron Z. For Mn 2RuGa, they satisfy\nMtot=Z−24, and the number of valence electron is 25,TABLE I: The SHG susceptibilities for different materials.\n|χ(2)|and|χ(2)|SPrepresent the absolute values of SHG sus-\nceptibilities without and with spin polarization, respect ively,\nin units of pm/V. The unit of the photon energy is eV.\ncMQWs is the abbreviation of the coupled metallic quantum\nwells.\nMaterial |χ(2)| |χ(2)|SPPhoton energy Reference\nMn2RuGa 3614 0.38 This work\nGaAs 358 1.80 This work\nGaAs 350 1.53 Ref.[35]\nTaAs 3600 1.55 Ref.[36]\nCo3Sn2S21050.05 Ref.[37]\ncMQWs 1500 1.35 Ref.[38]\nBiFeO 315-19 0.80 Ref.[39]\nCaCoSO 6.9 1.17 Ref.[40]\nsoMtotis about 1µB[34].\nIn orderto understand the nonlinear optical properties\nof Mn 2RuGa, we have calculated the second-order non-\nlinear optical susceptibilities. Nonmagnetic Mn 2RuGa\nbelongs to the point group T d. There are six equiva-\nlent nonvanishing SHG susceptibilities χ(2)\nxyz=χ(2)\nxzy=\nχ(2)\nyxz=χ(2)\nyzx=χ(2)\nzxy=χ(2)\nzyx. When the antiferromag-\nnetic coupling appears along the zdirection, the symme-\ntry is reduced from 24 to 8, belonging to the magnetic\npoint group -42m. This changes SHG. The system con-\ntains three independent nonvanishing elements, namely,\nχ(2)\nxyz=χ(2)\nyxz,χ(2)\nxzy=χ(2)\nyzx, andχ(2)\nzxy=χ(2)\nzyx. The SHG\nsusceptibility satisfies the intrinsic permutation symme-\ntry, that is χ(2)\nabc=χ(2)\nacb. Thus, there are only two inde-\npendent nonvanishing elements, χ(2)\nxyz=χ(2)\nxzy=χ(2)\nyxz=\nχ(2)\nyzx, andχ(2)\nzxy=χ(2)\nzyx.\nFor the SHG susceptibility χ(2)\nxyz, the absolute value of\nthe spin-up channel is given by the red solid line in Fig.\n1(b). It is shown that |χ(2)\nxyz|is at the maximum value\nof 11670.76 pm/V when the photon energy approaches 0\neV. Due to the metallic nature for the spin-up channel\nin Mn 2RuGa, the intensity of |χ(2)\nxyz|at 0 eV is not accu-\nrate, which also depends on the damping parameter [see\nAppendix A for more details]. When the photon energy\nis between 0 and 0.22 eV, |χ(2)\nxyz|decreasesmonotonically.\nAt 0.22 eV, the value of |χ(2)\nxyz|is close to zero. As the\nphoton energyincreases, a dramaticpeak appearsat 0.38\neV, as shown in Table I. The intensity of this peak is as\nlarge as 3614.37 pm/V. We also note that this peak is in-\nsensitive to the damping parameter, as discussed in Ap-\npendix A. As the energy further increases, the spectrum\noscillates and gradually decreases, and the final intensity\nis close to zero. By contrast, the spin-down |χ(2)\nxyz|[see\nthe black dashed line in Fig. 1(b)] is much smaller, with\na maximum of about 652.46 pm/V in the entire energy\nrange. This is consistent with the experimental finding\n[41], where only the majority spin channel is optically4\nexcited highly. When the energy is less than 0.2 eV, the\nintensity difference of |χ(2)\nxyz|between the spin-down and\nspin-up channels is largest. In the energy range of 0.2\nto 1.6 eV, the spin-down |χ(2)\nxyz|change is relatively sta-\nble, but the spin-up decreases, so the difference between\nthe two spins decreases. When the energy is larger than\n1.6 eV, the spin-down |χ(2)\nxyz|is very close to the spin-up,\nand the intensity is almost zero. Therefore, we can con-\nclude that SHG in Mn 2RuGa is mainly contributed by\nthe spin-up channel [41], while the contribution of spin-\ndown channel is negligible. As a result, we obtain nearly\n100% spin-polarized SHG in Mn 2RuGa. We also notice\nthat GaAs and Mn 2RuGa share the same point group.\nHowever, as shown in Fig. 1(c), the maximum intensity\nof|χ(2)\nxyz|in GaAs is only about 358.19 pm/V at 1.8 eV,\nwhich is about 10 times smaller than the spin-polarized\nSHG of Mn 2RuGa. The SHG susceptibilities for other\nmaterials are also presented in Table I. We believe that\nsuch a large spin-polarized SHG in Mn 2RuGa has poten-\ntial applications in spin-filter devices [42].\nB. Intraband and interband contributions\nTo reveal insights into this nearly 100% spin-polarized\nSHG, we resort to the band structure of Mn 2RuGa, as\nshown in Fig. 2(a). The red line indicates the spin-up\nchannel, and the black line indicates the spin-down chan-\nnel. The band of spin-up channel crosses the Fermi level\nat multiple kpoints, indicating a metallic state. This re-\nsultcanbeverifiedfromthetotaldensityofstates(DOS),\nas shown in Fig. 2(b). Bands in the energy range from\n−0.08 to 0.12 eV are mainly occupied by dorbitals, in\nwhich Mn-3 dorbitals are dominant, Ru-4 dorbitals con-\ntribute less, and the contribution from Ga-3 dorbitals\nis negligible. However, the spin-down band only touches\nthe Fermi level near the Γ point. The total DOS near the\nFermi level is close to 0. Therefore, Mn 2RuGa is a half\nmetal, which is consistent with previous report [29, 30].\nThe band gap disappears in the spin-up channel, but ap-\npears in the spin-down channel. Therefore, the appear-\nance of band gap hinders the transition of electrons from\nthe valence band to the conduction band.\nNext, we further analyze the difference between the\nspin-up and the spin-down χ(2)\nxyzof Mn 2RuGa by exam-\nining the real and imaginary parts separately. For the\nspin-up channel, the real and imaginary parts of χ(2)\nxyzare\ngiven in red solid and black dashed lines, respectively, as\nshown in Fig. 2(c). The real part of χ(2)\nxyz(Re(χ(2)\nxyz))\ndecreases monotonously as the photon energy increases\nfrom 0 to 0.16 eV. When the photon energy reaches 0.16\neV,itsvaluereachesaminimum of-1099.71pm/V.When\nthe energy is larger than 0.16 eV, the spectrum oscillates\nand finally approaches zero. For the imaginary part of\nχ(2)\nxyz, that is Im( χ(2)\nxyz), a Lorentzian-like resonance ap-\npears in the energy range between 0 and 0.4 eV. Com-\npared with the real part, its first negative peak shiftsΓ L K0.01.0\n-0.5\n-1.0W \b X0.5\n0.03.06.09.0\n-3.0\n-6.0\n-9.0\n0.0 0.5 1.0 1.5 2.0 -0.5-1.0-1.5-2.0DOS (States/eV)\nE (eV) E (eV)(a) (b)Total Mn-3d Ru-4d\nSpin-up Spin-dnW1\nW2\nW3\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(c)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) \n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (d)\n(e) (f)Spin-up Spin-up\nSpin-dn Spin-dnInter (\t) Intra (\n)\nInter (2\u000b) Intra (2\f)\nInter (\r) Intra (\u000e)\nInter (2\u000f) Intra (2\u0010)\n05001000\n-500\n-10001500\n-150004000800012000\n-4000\n-8000\n-12000\nIm(χxyz) (pm/V)(2)Im(χxyz) (pm/V)(2)χxyz (pm/V)(2)χxyz (pm/V)(2)04000800012000\n-4000\n-8000\n-12000\n05001000\n-500\n-10001500\n-1500\nFIG.2: (a)BandstructurewithoutSOCforMn 2RuGa, where\nred line represents the spin-up channel, and black line repr e-\nsents the spin-down channel. The blue circles represent the\ndxy,dyzanddxzorbitals of Mn atoms for the bands W 1, W2\nand W 3. The dashed line denotes the Fermi level. (b) To-\ntal DOS of Mn 2RuGa, where the partial DOS of Mn-3 dand\nRu-4dstates are also given and represented by red and green\nlines, respectively. The upward arrow indicates the spin-u p\nchannel, and the downward arrow indicates the spin-down\nchannel. The vertical dashed line denotes the Fermi level. ( c)\nReal and imaginary parts of the SHG susceptibility χ(2)\nxyzfrom\nthe spin-up channel in Mn 2RuGa without SOC. (d) Calcu-\nlated Im( χ(2)\nxyz) from inter( ω)/(2ω) (black solid, black dashed\ncurve)andintra( ω)/(2ω)(red solid, reddashed-dottedcurve)\nparts. (e) and (f) are similar to (c) and (d), but from the spin -\ndown channel.\nto the lower energy by about 0.06 eV, and the intensity\nis decreased to 7768.47 pm/V. When the photon energy\nfurther increases, the spectrum oscillates.\nIn general, the imaginary part of χ(2)\nxyzreflects the op-\ntical absorption in NLO experiments. From the band\nstructure, the optical absorption involves in the intra-\nand interband currents. Thus, we decompose Im( χ(2)\nxyz)\ninto the inter- and intraband parts for the spin-up chan-\nnel, asshowninFig. 2(d). It clearlyshowsthattheintra-\nband contribution dominates the spectrum in the lower\nphoton energy window from 0 to 0.4 eV. By contrast,\nthe interband contribution from both single- and two-\nphoton resonances is much smaller. In the energy range\nof around 0.4 −2.0 eV, these four spectra are comparable\nand have opposite signs for the single- and two-photon\nresonances, leading to the oscillation of Im( χ(2)\nxyz) in this5\nenergy range. Thus, the negative characteristic peak of\nIm(χ(2)\nxyz) in the lower energy is determined by the in-\ntraband current. In other words, the interband current\ncontributes little to the negative characteristic peak in\nthe lower energy. We know that the negative character-\nistic peak locates in the energy range of 0 −0.4 eV. The\ntwo-photon resonance would correspond to the energy\nrange of around 0 −0.8 eV. In this perspective, the bands\nrelated to the intraband current would locate in the en-\nergy range from −0.4 to 0.4 eV near the Fermi level. As\nshown in Fig. 2(a), we can see that only three bands W 1,\nW2and W 3appear in this energy range. The bands W 1\nand W 2are degenerate along the L-Γ direction, while\nthe bands W 2and W 3are degenerate along the Γ- Xdi-\nrection. More importantly, these three bands disperse\nquadratically along the L-Γ-Xdirection. This means\nthat they highly disperse along this high-symmetry line,\ncontributingalargenormalvelocitytothe intrabandcur-\nrent [43]. This is the reason why the intraband current\ndominates the negative characteristic peak in the lower\nenergy range.\nWealsonoticethatthereexisttworegionsfortheinter-\nbandtransitionsamongthesethreebands. Oneislocated\nnear theLpoint along the L-Γ direction, where the dou-\nble degenerate bands of W 1and W 2form the conduction\nbands while the band W 3is the valence band. The other\ninterband transition appears near the Xpoint along the\nΓ-Xdirection,wherethebandW 1istheconductionband\nwhile the two-fold degenerate bands of W 2and W 3are\nthe valence bands. According to the selection rules, the\ninterband transitions from these two regions are largely\nlimited. This is because these three bands are mainly\nformed by the dxy,dyzanddxzorbitals of Mn atoms\n[see Fig. 2(a) for more details]. The interband transi-\ntions between the same dorbitals are not allowed. As\na result, the interband current contributes little to the\nnegative characteristic peak in the lower energy range.\nThis is generic for metallic ferro- or ferrimagnets as the\nspin-splitting states near the Fermi level are dominated\nby thedorbitals.\nIn the case of spin-down channel, the real and imag-\ninary parts of χ(2)\nxyzare comparable, as shown in Fig.\n2(e). Both oscillate around zero as the photon energy\nincreases. However, they are largely suppressed in com-\nparison with those of spin-up channel. Similarly, we also\ndecompose Im( χ(2)\nxyz) into the inter- and intraband con-\ntributions, as shown in Fig. 2(f). It is found that both\nthe inter- and intraband currents are comparable and os-\ncillate around zero, which are much smaller than those\nof the spin-up channel. This is because only few elec-\ntrons are allowed to transit from the valence band to the\nconduction band near the Γ point. However, the large\nband gap of the spin-down channel limits both the inter-\nand intraband currents. This explains why SHG of the\nspin-down channel is much smaller.C. SHG in metallic Mn 2Ga\nTo further confirm the contribution of SHG from the\nspin-up quadratic bands in Mn 2RuGa, we artificially re-\nmove the Ru atoms in Mn 2RuGa and obtain the crystal\nstructure of Mn 2Ga [30]. In Mn 2Ga, two Mn atoms are\nantiferromagnetic coupled and their magnetic moments\nare close to 3 µB, which can compensate each other. The\nmagnetic moments of Ga atoms are very small, so the to-\ntal magnetic moment of the unit cell is nearly zero. This\ncoincides with the previous report [34]. However, the\nantiferromagnetic coupling in Mn 2Ga has a huge effect\non the spin-polarized band structures, as shown in Fig.\n3(a). It is found that both the spin-up and spin-down\nbands cross the Fermi level, indicating that Mn 2Ga is a\nmetal. This is also confirmed from the PDOS, as shown\nin Fig. 3(b). We can see that obvious DOS exits both\nfor the spin-up and spin-down channels near the Fermi\nlevel, where the Mn-3 dorbitals dominates.\nΓ L K0.02.0\n-1.00.5\n-2.0\nW W\u0011-1.5-0.51.01.5\nSpin-up Spin-dnE (eV)(a)\n0.0 0.5 1.0 1.5 2.0 2.5 3.050001000015000\n0\nE (eV) (c)0.03.0\n6 \u0012 \u00139 \u0014 \u0015-3.0- \u0016 \u0017 \u0018\u0019 \u001a \u001b \u001c\n0.0 0.5 1.0 1.5 2.0 -0.5-1.0-1.5-2.0DOS (States/eV)\nE (eV) (b)\n(d)Total Mn-3d\nSpin-dn\nInter (\u001d) Intra (\u001e)\nInter (2\u001f) Intra (2 )\nx ! \" # $ % & ' ( ) * + , . /\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) 03000\n-30000 1 2 3 45 7 8 : ;Im(\n?@ABCD(2)Spin-up\nSpin-dn\n-12000\nFIG. 3: (a) Band structures for Mn 2Ga, where the red and\nblack lines represent the spin-up and spin-down channels, r e-\nspectively. The dashed line denotes the Fermi level. (b) Tot al\nDOS of Mn 2Ga, where the partial DOS of Mn-3 dorbitals is\nalso given and represented by red line. The upward arrow\nindicates the spin-up channel, and the downward arrow indi-\ncates the spin-down channel. The vertical dashed line denot es\nthe Fermi level. (c) The absolute value of the SHG suscep-\ntibilityχ(2)\nabcfor Mn 2Ga, where abcrefers to xyz,xzy,yxz,\nandyzx. The red solid and black dashed lines denote the\nspin-up and spin-down channels, respectively. (d) Calcula ted\nIm(χ(2)\nxyz) from inter( ω)/(2ω)(black solid, black dashed curve)\nand intra ( ω)/(2ω) (red solid, red dashed-dotted curve) parts\nfor the spin-down channel.\nFor Mn 2Ga, SHG has two independent nonvanishing\nelements, namely, χ(2)\nxyz=χ(2)\nxzy=χ(2)\nyxz=χ(2)\nyzx, and\nχ(2)\nzxy=χ(2)\nzyx. The absolute value of χ(2)\nxyzis displayed\nin Fig. 3(c). For the spin-up channel, the first peak\nappears at the photon energy 0.14 eV and its intensity\nis 5602.48 pm/V. Then, the intensity sharply decreases\nand finally approaches zero as the photon energy further6\nincreases. In the case of the spin-down channel, χ(2)\nxyz\nis similar to that of the spin-up channel, but the inten-\nsity is much larger. However, this is contrast to that of\nthe spin-down channel in Mn 2RuGa, where χ(2)\nxyzis much\nsmaller. This is because no band gapappears in the spin-\ndown channel of Mn 2Ga and three metallic bands cross\nthe Fermi level. More importantly, one of them disperses\nquadratically with knear the Γ point, which largely con-\ntributes to the intraband part of χ(2)\nxyz. This implies that\nthe presence of band gap in the spin-down channel of\nMn2RuGa limits χ(2)\nxyz, while the absence of band gap or\nthe quadratic band near the Fermi level enhances χ(2)\nxyz.\nTo this end, we decompose Im( χ(2)\nxyz) of the spin-down\nchannel into the inter- and intraband contributions, as\nshown in Fig. 3(d). It clearly shows that the intraband\ncurrent dominates, while the interband contribution is\nnearly neglected. In addition, the two-photon resonance\nis obviously larger than that of one-photon resonance.\nThis would be easily detected by SHG in experiment.\nD. Role of SOC in SHG in Mn 2RuGa and the\ngroup symmetry\nWith SOC, the spin-up and the spin-down channels\nmix together. The band structure of Mn 2RuGa is shown\nin Fig. 4(a), which almost coincides with the spin-\npolarized band structures. Figure 4(b) shows that DOS\nhas a peak between −0.1 and 0.1 eV, and mainly comes\nfrom the Mn-3 dorbitals, which is the same as the spin-\nup channel. This is because the spin-down DOS is close\nto zero in this energy range. However, in the energy\nrange below −0.1 eV and above 0.1 eV, DOS under SOC\nchanges compared to the spin-up channel, which is due\nto the contribution of the spin-down channel. This will\naffect SHG.\nIn fact, if we only include SOC, but ignore the mag-\nnetic field direction, the symmetries of the system re-\nmain unchanged. In real calculations, SOC is considered\nthrough a tiny magnetic field applied along the z-axis.\nThe symmetry is reduced and belongs to the magnetic\npoint group of -4. The remaining four symmetry op-\nerations are the identity operation E, the twofold rota-\ntional symmetry C2zwith the binary axis as the zaxis,\nand two combination operations IC4zandIC−1\n4z.IC4z\ndenotes the rotation of π/2 around the z-axis, followed\nby a mirror symmetry σxy.IC−1\n4zis similar to IC4z,\nbut with a rotation −π/2 around the z-axis. As a re-\nsult, there are six independent nonvanishing elements,\nχ(2)\nxyz=χ(2)\nyxz,χ(2)\nxzy=χ(2)\nyzx,χ(2)\nzxy=χ(2)\nzyx,χ(2)\nxxz=−χ(2)\nyyz,\nχ(2)\nxzx=−χ(2)\nyzy, andχ(2)\nzxx=−χ(2)\nzyy. Based on the in-\ntrinsic permutation symmetry, only four components in-\ndeed appear, χ(2)\nxyz=χ(2)\nxzy=χ(2)\nyxz=χ(2)\nyzx,χ(2)\nzxy=χ(2)\nzyx,\nχ(2)\nxxz=χ(2)\nxzx=−χ(2)\nyyz=−χ(2)\nyzy, andχ(2)\nzxx=−χ(2)\nzyy. It\nshould be noted that χ(2)\nxxzandχ(2)\nzxxare induced from the\nreduced magnetic point group.Γ L K0.01.0\n-0.5\n-1.0\nW W X0.5E (eV)(a)\n0.03.0\nE F GH I J0.0 0.5 1.0 1.5 2.0 -0.5-1.0-1.5-2.0DOS (States/eV)\nE (eV) (b)\nTotal Mn-3d Ru-4d\n0.0 0.5 1.0 1.5 2.0 2.5 3.03000\nK L M NO P Q R0\nE (eV) (c)S T U V Y Z [ \\ ] ^ _ ` a b c\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (d)d e f g h i j k l m n o p q r s tz u v w y { } ~ \n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(e)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (f)\nInter ( ) Intra ( )\nInter (2 ) Intra (2 )020004000\n \n-2000\n-4000 020004000\n \n-2000\n-4000 \n abc\n¡¢£¤(2)¥¦ abc\n§¨©ª«¬(2) Im(\n®xxz) (pm/V)(2)¯xxz (pm/V)(2)3000\n° ± ² ³´ µ ¶ ·0\nFIG. 4: (a) Band structure for Mn 2RuGa under SOC. The\ndashed line denotes the Fermi level. The rectangles denote\nthe flat valence and conduction bands and the arrow labels\nthe transitions between those flat bands. (b) Total DOS of\nMn2RuGa, where the partial DOS of Mn-3 dand Ru-4 dstates\nare also given and represented by red and green lines, re-\nspectively. The vertical dashed line denotes the Fermi leve l.\n(c) and (d) The absolute values of the SHG susceptibility\nχ(2)\nabcfor Mn 2RuGa under SOC. (e) Real and imaginary parts\nofχ(2)\nxxz. (f) Calculated Im( χ(2)\nxxz) from inter( ω)/(2ω) (black\nsolid, black dashed curve) and intra ( ω)/(2ω) (red solid, red\ndashed-dotted curve) parts.\nFigure 4(c) shows the absolute values of SHG suscep-\ntibilitiesχ(2)\nxyzandχ(2)\nzxy. It clearly shows that those two\nSHG spectra are nearly the same as that of the spin-\nup channel [see Fig. 1(d)]. This is because the contri-\nbution from the spin-down channel is negligible. Thus,\nSOC has little effect on these six SHG spectra, which\nalso appear in the spin-polarized case. It is also found\nthat the vanishing SHG spectra χ(2)\nxxzandχ(2)\nzxxin the\nspin-polarized case are recovered now, as shown in Fig.\n4(d). The intensity of χ(2)\nzxxis obviously larger than that\nofχ(2)\nxxz. The appearance ofthem directly comes from the\nreduced magnetic point group. In the following, we use\nthe four remaining symmetries to understand the nonva-\nnishing SHG spectrum χ(2)\nxxz. The matrix representations\nofEandC2zare diag{1,1,1}and diag {−1,−1,1}. The7\nother twoIC4zandIC−1\n4zare\nIC4z=\n0 1 0\n−1 0 0\n0 0 −1\n,IC−1\n4z=\n0−1 0\n1 0 0\n0 0 −1\n.\n(9)\nThe transformations of position operator under these\nfour symmetry operations are as follows. E: (x,y,z)→\n(x,y,z),C2z: (x,y,z)→(−x,−y,z),IC4z: (x,y,z)→\n(y,−x,−z), andIC−1\n4z: (x,y,z)→(−y,x,−z). As a\nresult, we can get:\nE:χ(2)\nxxz→χ(2)\nxxz,\nC2z:χ(2)\nxxz→χ(2)\n(−x)(−x)z=χ(2)\nxxz,\nIC4z:χ(2)\nxxz→χ(2)\nyy(−z)=−χ(2)\nyyz,\nIC−1\n4z:χ(2)\nxxz→χ(2)\n(−y)(−y)(−z)=−χ(2)\nyyz.(10)\nIt clearly shows that these four SHG susceptibilities do\nnot cancel out each other. We can obtain these induced\nSHG spectra via the above symmetry analysis, where\nχ(2)\nxxzandχ(2)\n−yyzare protected by C2zandIC4z, respec-\ntively.\nThe reduced magnetic point group is closely related\nto SOC and the applied magnetic filed direction. This\nmeans that the induced SHG spectra must have a deep\nrelation to them. Here, we take χ(2)\nxxzto reveal the un-\nderlying physics. The decomposed real and imaginary\nparts ofχ(2)\nxxzare displayed in Fig. 4(e). We can see\nthat both have a similar manner. Sizable intensities are\nmainly located in the photon energy range of 0 −0.75 eV.\nTo understand the contributions of inter- and intraband\ncurrents, we decompose Im( χ(2)\nxxz) into the inter- and in-\ntraband parts, as shown in Fig. 4(f). There also exists a\nlarge contribution from the intraband current, where the\nhighly dispersed bands near the Fermi level play a role.\nIt should be noted that the interband contributions are\nlargely enhanced, which are nearly limited in the spin-\npolarized case. We find that many transitions are now\nallowed between the flat valence and conduction bands,\nas shown in rectangles of Fig. 4(a). However, those tran-\nsitions are not allowed in the spin-polarized case. This is\nbecause the flat valence bands are originally spin-up po-\nlarized, while those flat conduction bands are spin-down\npolarized. The direct transition from the spin-up band\nto the spin-down band is forbidden. But, this does occur\nunder SOC as the conservation of spin is not needed. To\ncheck it, we calculate some matrix elements of the posi-\ntionoperatorbetweenthoseflatbandsatseveral kpoints,\nas shown in Table II. We can see that the y-componentof\nmatrix elements are small but not zero. However, the x-\nandz-components are much larger. The maximum abso-\nlute value reaches as large as 8.02 a 0with a 0being the\nBohrradius. Thesenonzeromatrixelementsconfirmthat\nthe transitions are not allowed in the spin-polarized case,\nbut did occur with SOC. This tells us that the reduced\nmagnetic point group or the remaining four symmetriesprotect the induced SHG spectra, which is similar to our\nprevious study [44].\nTABLE II: The matrix elements of position operator between\nthose flat bands near the Fermi level under SOC in Mn 2RuGa\nfor three kpoints, where the atomic unit is used.\nkpointx y z\nRe Im Re Im Re Im\n(0,0.139,0) -0.83 0.32 0.05 0.12 -0.68 -1.74\n(0,0.209,0) -1.07 -2.34 0.01 -0.01 -2.49 1.13\n(0,0.278,0) 1.93 2.45 -0.09 0.07 6.29 -4.96\nIV. CONCLUSIONS\nIn conclusion, we have demonstrated that large nearly\n100% spin-polarized SHG carries rich information about\nthe electronic structures of the half-metal Mn 2RuGa.\nThe band gap in the spin-down channel limits SHG.\nIn contrast, the spin-up channel is metallic and gives\nrise toχ(2)\nxyzas large as 3614 pm/V, which is about 10\ntimes largerthan that of typical nonlinear materials such\nas GaAs. For the spin-up channel, the intraband cur-\nrent mainly contributes to χ(2)\nxyz, which stems from three\nhighly dispersed bands near the Fermi level. In addition,\nSOC under the z-axis magnetization field induces addi-\ntionalSHG susceptibilities suchas χ(2)\nxxzfrom the reduced\nmagnetic point group, where the interband transitions\ndominate. Our study would provide a good guide in fu-\nture application of largespin-polarized SHG in spin-filter\ndevices.\nACKNOWLEDGMENTS\nThis work was supported by the National Science\nFoundationofChinaunderGrantNo. 11874189. Wealso\nacknowledge the Fermi cluster at Lanzhou University for\nproviding computational resources. GPZ was supported\nby the U.S. Department of Energy under Contract No.\nDE-FG02-06ER46304. The research used resources of\nthe National Energy Research Scientific Computing Cen-\nter,whichissupportedbytheOfficeofScienceoftheU.S.\nDepartment of Energy under Contract No. DE-AC02-\n05CH11231.\n∗sims@lzu.edu.cn\n†guo-ping.zhang@outlook.com\nAPPENDIX A: The crystal structure of Mn 2RuGa\nand the effect of damping parameters on SHG\nHeusler alloyshavethree different structures belonging\nto different space groups [45]. The normal full-Heusler\nX2YZalloys belong to group symmetry L2 1(No. 225).8\n0.0 0.5 1.0 1.5 2.0 2.5 3.050001000015000\n0\nE (eV) Spin-up(b)\n|χxyz| (pm/V)(2)η¸ ¹ º » ¼ ½ ¾¿ À Á  à Ä\nηÅ Æ Ç È É Ê ËÌ Í Î Ï Ð Ñ\nη Ò Ó Ô Õ Ö × ØÙ Ú Û Ü Ý Þ200002500030000\nz\nxyMn1\nMn2Ruß à(a)\nFIG. 5: (a) Crystal structure of Mn 2RuGa. Purple, or-\nange, grey, and green spheres represent Mn 1, Mn2, Ru, and\nGa atoms, respectively. The arrows marked on the Mn atoms\nindicate the local magnetic moments, forming the antiferro -\nmagnetic configuration. (b) The absolute value of the spin-u p\nχ(2)\nxyzof Mn 2RuGa without SOC for different damping param-\netersη.\nThe Half-Heusler XYZcompounds have group symme-\ntryC1b(No. 216), and the inverse-Heusler X2YZalloys\nwith group symmetry XA(No. 216) [28]. We select\na stable one, that is XAstructure, as our example, as\nshown in Fig. 5(a), where the sublattice Mn 2RuGa are\nferrimagnetic ordering. The magnetic moments of these\nsublattice Mn atoms are 3.13 and -2.29 µB, respectively.\nAs aresult, the net magneticmoment ofunit cellisabout\n1µB, which agrees well with the Slater-Pauling rule.\nDue to the metallic nature for the spin-up channel in\nMn2RuGa, the nonlinear optical response at zero fre-\nquency must be estimated, because the energy difference\n¯hωnm=En−Emis zero for those metallic states near\nthe Fermi level. As a result, ωnm−ωorωnm−2ωin\nthe denominators of Eqs. (4)-(7) in the main text di-\nverges. To avoid this, the damping parameter ηis in-\ntroduced to estimate |χ(2)|at zero frequency. However,\nthe obtained results are still not accurate. As shown in\nFig. 5(b), we can see that the intensity of the spin-up\n|χ(2)\nxyz|at 0 eV largely depends on η. Whenη= 0.004\nHartree, the intensity of spin-up |χ(2)\nxyz|at 0 eV is about\n6150.37 pm/V. When ηis decreased to 0.003 Hartree,\nthe value is about twice larger than that for η= 0.004\nHartree. When ηis further decreased to 0.002 Hartree,\nthe intensity of spin-up |χ(2)\nxyz|is dramatically increased.\nThis means that |χ(2)|usually diverges at zero frequency.\nIn other words, it is a challenge to accurately compute\n|χ(2)|at zero frequency. This is also the case of linear\nresponse in Hall effect. According to the Drude model,\nthe frequency-dependent conductivity σ(ω) is given by,\nσ(ω) =σ0\n1−iωτ, (11)\nwhereσ0is the DC Drude conductivity without the ex-\nternal magnetic field and τis the relaxation time. At\nzero frequency, we can see σ(ω) reduces to σ0=ne2τ/m\nwithn,e, andmbeing the electron density, the electron\ncharge, and the electron mass, respectively. This shows\nthe failure of zero-frequency response already exists in\nthe linear response, which would be our future researchfocus of SHG.\nWhen the photon energy is further increased to about\n0.38 eV, a stable peak of spin-up |χ(2)\nxyz|with respect to\nηappears. The intensity of this peak increases as η\ndecreases. But the change is small. This implies that\nthe intensity of this second peak is insensitive to η. We\nalso note that the intensity of this peak corresponding to\nη= 0.003Hartree is as large as3614.37pm/V, which can\nbe compared with SHG spectra of other materials such\nas GaAs, TaAs, and CaCoSO [see Table I].\nAPPENDIX B: The SHG susceptibility for\nMn2RuGa without SOC\n0.0 0.5 1.0 1.5 2.0 2.5 3.030006000900012000\n0\nE (eV)Spin-up\nSpin-dn\ná â ã ä å æ çèχabc\néêëìíîï(2)\nFIG. 6: The absolute value of the SHG susceptibilities χ(2)\nabc\nfor Mn 2RuGa without SOC. The red solid and black dashed\nlines denote the spin-up and spin-down channels, respectiv ely.\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(a)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) \n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (b)\n(c) (d)Spin-up Spin-up\nSpin-dn Spin-dnInter (ω) Intra (ω)\nInter (2ω) Intra (2ω)\nInter (ω) Intra (ω)\nInter (2ω) Intra (2ω)\n05001000\n-500\n-10001500\n-150004000800012000\n-4000\n-8000\n-12000\nIm(χzxy) (pm/V)(2)Im(χzxy) (pm/V)(2)χzxy (pm/V)(2)χzxy (pm/V)(2)04000800012000\n-4000\n-8000\n-12000\n05001000\n-500\n-10001500\n-1500\nFIG. 7: (a) The real and imaginary parts of the SHG sus-\nceptibility χ(2)\nzxyfrom the spin-up channel in Mn 2RuGa with-\nout SOC. (b) Calculated Im( χ(2)\nzxy) from inter( ω)/(2ω) (black\nsolid, black dashed curve) and intra ( ω)/(2ω) (red solid, red\ndashed-dotted curve) parts. (c) and (d) are similar to (a) an d\n(b), but from the spin-down channel.\nNext, theabsolutevalueofthe SHGsusceptibility χ(2)\nzxy\nfor Mn 2RuGa without SOC is given in Fig. 6. For the\nspin-upchannel,theintensityatzerofrequencyis9936.41\npm/V, which is smallerthan that of χ(2)\nxyz. The secondary9\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(c)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (d)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(e)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (f)0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) Re\nIm(a)\n0.0 0.5 1.0 1.5 2.0 2.5 3.0\nE (eV) (b)Inter (ω) Intra (ω)\nInter (2ω) Intra (2ω)\nInter (ω) Intra (ω)\nInter (2ω) Intra (2ω)\nInter (ω) Intra (ω)\nInter (2ω) Intra (2ω)4000800012000\n0\n-8000\n-12000-4000χxyz (pm/V)(2)χzxy (pm/V)(2)χzxx (pm/V)(2)\nIm(χxyz) (pm/V)(2)Im(χzxy) (pm/V)(2)Im(χzxx) (pm/V)(2)4000800012000\n0\n-8000\n-12000-4000\n4000800012000\n0\n-8000\n-12000-4000\n4000800012000\n0\n-8000\n-12000-40004000800012000\n0\n-8000\n-12000-4000\n4000800012000\n0\n-8000\n-12000-4000\nFIG. 8: (a) The real and imaginary parts of χ(2)\nxyz. (b) Calcu-\nlated Im( χ(2)\nxyz) from inter( ω)/(2ω) (black solid, black dashed\ncurve)andintra( ω)/(2ω)(redsolid, reddashed-dottedcurve)\nparts. (c) and (d) are similar to (a) and (b), but for χ(2)\nzxy. (e)\nand (f) are similar to (a) and (b), but for χ(2)\nzxx.\npeak appears at 0.41 eV with an intensity of 3452.36\npm/V. When the photon energy is larger than 0.59 eV,the spectrum oscillates and then approaches to 0. In\ncontrast,the intensityofthespin-down χ(2)\nzxyisverysmall\nin the energy range from 0 to 3 eV. Therefore, in the low\nenergy region, the intensity of the spin-up χ(2)\nzxyis much\nstronger than that of the spin-down channel, which is\nsimilar toχ(2)\nxyz. This is because the spin-up channel has\nno band gap, while a gap in the spin-down channel limits\nSHG.\nThe real and imaginary parts of the spin-up channel\nforχ(2)\nzxyis shown in Fig. 7(a). We decompose Im χ(2)\nxyz\nof the spin-up channel into the inter- and intraband con-\ntributions, as shown in Fig. 7(b). It is found that the\nintraband contributions are dominant in the low energy\nregion. For the spin-down channel, the real and imag-\ninary parts are very small, as shown in Fig. 7(c). It\nmainly comes from the intraband and interband tran-\nsitions, as shown in Fig. 7(d). These results of both\nspin-up and spin-down channels are similar to χ(2)\nxyz.\nAPPENDIX C: The SHG susceptibility for\nMn2RuGa with SOC\nWith SOC, χ(2)\nxyzandχ(2)\nzxyare similar to the spin-up\nχ(2)\nxyzandχ(2)\nzxy, respectively, as shown in Figs. 8(a)-8(d).\nIt shows that SOC has little effect on these two compo-\nnents. On the contrary, the reduced symmetries induce\nχ(2)\nxxzandχ(2)\nzxx, showing different characteristics. The\nreal and imaginary parts of χ(2)\nzxxare shown in the red\nand black lines in Fig. 8(e). It is contributed by inter-\nand intraband transitions, as shown in Fig. 8(f). Com-\npared with the spin-polarized χ(2)\nxyzandχ(2)\nzxy, the inter-\nband contribution increases, which is similar to χ(2)\nxxz.\n[1] Y. R. Shen, The Principles of Nonlinear Optics (Wiley,\nNew York, 1984).\n[2] R. W. Boyd, Nonlinear Optics (Elsevier Science, Ams-\nterdam, 2003).\n[3] P.Franken, A.E.Hill, C. Peters, andG. Weinreich, Phys.\nRev. Lett. 7, 118 (1961).\n[4] M.W.Dohertya, N.B. Mansonb, P.Delaneyc, F.Jelezko,\nJ. Wrachtrupe, and L. C. L. Hollenberg, Phys. Rep. 528,\n1 (2013).\n[5] M. L. Goldman, A. Sipahigil, M. W. Doherty, N. Y. Yao,\nS. D. Bennett, M. Markham, D. J. Twitchen, N. B. Man-\nson, A. Kubanek, andM. D. Lukin, Phys. Rev.Lett. 114,\n145502 (2015).\n[6] A. Abulikemu, Y. Kainuma, T. An, and M. Hase, ACS\nPhotonics 8, 988 (2021).\n[7] L. Jia, Y. K. Song, J. L. Yao, M. S. Si, and G. P. Zhang,\nPhys. Rev. B 105, 214309 (2022).\n[8] Y. L. Li, Y. Rao, K. F. Mak, Y. You, S. Wang, C. R.\nDean, and T. F. Heinz, Nano Lett. 13, 3329 (2013).\n[9] M. Gr¨ uning and C. Attaccalite, Phys. Rev. B 89,\n081102(R) (2014).\n[10] S. H. Rhim, Y. S. Kim, and A. J. Freeman, Appl. Phys.\nLett.107, 241908 (2015).[11] Z. Y. Sun, Y. F. Yi, T. C. Song, G. Clark, B. Huang, Y.\nW. Shan, S. Wu, D. Huang, C. L. Gao, Z. H. Chen, M.\nMcGuire, T. Cao, D. Xiao, W. T. Liu, W. Yao, X. D.\nXu, and S. W. Wu, Nature 572, 497 (2019).\n[12] W. S. Song, R. X. Fei, L. H. Zhu, and L. Yang, Phys.\nRev. B102, 045411 (2020).\n[13] R. X. Fei, W. S. Song, and L. Yang, Phys. Rev. B 102,\n035440 (2020).\n[14] G. P. Zhang and Y. H. Bai, Phys. Rev. B 103, L100407\n(2021).\n[15] P. E. Bl¨ ochl, Phys. ReV. B 50, 17953(1994).\n[16] G. Kresse and D. Joubert, Phys. Rev. B 59, 1758(1999).\n[17] G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).\n[18] G. Kresse and J. Hafner, Phys. Rev. B 49, 14251 (1994).\n[19] G. Kresse, and J. Furthm¨ uller, Phys. Rev. B 54, 11169\n(1996).\n[20] G. Kresse and J. Furthm¨ uller, Comput. Mat. Sci. 6, 15\n(1996).\n[21] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.\nLett.77, 3865 (1996).\n[22] J. E. Sipe and E. Ghahramani, Phys. Rev. B 48, 11705\n(1993).\n[23] C. Aversa and J. E. Sipe, Phys. Rev. B 52, 14636 (1995).10\n[24] J. L. P. Hughes and J. E. Sipe, Phys. Rev. B 53, 10751\n(1996).\n[25] B. Lu, S. Sayyad, M. ´A. S´ anchez-Mart´ ınez, K. Manna,\nC. Felser, A. G. Grushin, and D. H. Torchinsky, Phys.\nRev. Res. 4, L022022 (2022).\n[26] O. Rubel and P. Blaha, Computation 10, 22 (2022).\n[27] I. Galanakis and P. H. Dederichs, Phys. Rev. B 66,\n174429 (2002).\n[28] K. Fleischer, N. Thiyagarajah, Y. C. Lau, D. Betto, K.\nBorisov, C. C. Smith, I. V. Shvets, J. M. D. Coey, and\nK. Rode, Phys. Rev. B 90, 214420 (2014).\n[29] L. Wollmann, S. Chadov, J. K¨ ubler, and C. Felser, Phys.\nRev. B90, 214420 (2014).\n[30] G. P. Zhang, Y. H. Bai, M. S. Si, and T. F. George, Phys.\nRev. B105, 054431 (2022).\n[31] K. Fleischer, N. Thiyagarajah, Y. -C. Lau, D. Betto, K.\nBorisov, C. C. Smith, I. V. Shvets, J. M. D. Coey, and\nK. Rode, Phys. Rev. B 98, 134445 (2018).\n[32] H. Kurt, K. Rode, P. Stamenov, M. Venkatesan, Y. -C.\nLau, E. Fonda, and J. M. D. Coey, Phys. Rev. Lett. 112,\n027201 (2014).\n[33] S. Skaftouros, K. ¨Ozdo˘ gan, E. S ¸a¸ sıo˘ glu, and I. Galanakis,\nPhys. Rev. B 87, 024420 (2013).\n[34] I. Galanakis, K. ¨Ozdo˘ gan, E. S ¸a¸ sıo˘ glu, and S. Bl¨ ugel, J.\nAppl. Phys. 116, 033903 (2014).\n[35] S. Bergfeld and W. Daum, Phys. Rev. Lett. 90, 036801\n(2003).\n[36] L.Wu, S.Patankar, T.Morimoto, N.L.Nair, E.Thewalt,\nA. Little, J. G. Analytis, J. E. Moore, and J. Orenstein,\nNat. Phys. 13, 350 (2017).\n[37] K. Takasan, T. Morimoto, J. Orenstein, and J. E. Moore,\nPhys. Rev. B 104, L161202 (2021).\n[38] H. L. Qian, S. L. Li, C.-F. Chen, S. -W. Hsu, S. E. Bopp,\nQ. Ma, A. R. Tao and Z. W. Liu, Light Sci. Appl. 8, 13\n(2019).\n[39] R. C. Haislmaier, N. J. Podraza, S. Denev, A. Melville,\nD. G. Schlom, and V. Gopalan, Appl. Phys. Lett. 103,\n031906 (2013).\n[40] A. H. Reshak, Sci. Rep. 7, 46415 (2017).\n[41] C. Banerjee, N. Teichert, K. Siewierska, Z. Gercsi, G.\nAtcheson, P. Stamenov, K. Rode, J. M. D. Coey, and J.\nBesbas, Nat. Commun. 11, 4444 (2020).\n[42] W. -F. Tsai, C. -Y. Huang, T. -R. Chang, H. Lin, H. -T.\nJeng, A. Bansil, Nat. Commun. 4, 1500 (2013).\n[43] D. Xiao, M. C. Chang, and Q. Niu, Rev. Mod. Phys. 82,\n1959 (2010).\n[44] L. Jia, Z. Y. Zhang, D. Z. Yang, M. S. Si, G. P. Zhang,\nand Y. S. Liu, Phys. Rev. B 100, 125144 (2019).\n[45] M. Hakimi, M. Venkatesan, K. Rode, K. Ackland, and J.\nM. D. Coey, J. Appl. Phys. 113, 17101 (2013)." }, { "title": "1109.2705v2.Interplay_between_non_equilibrium_and_equilibrium_spin_torque_using_synthetic_ferrimagnets.pdf", "content": "arXiv:1109.2705v2 [cond-mat.mes-hall] 10 Apr 2012Interplay between non equilibrium and equilibrium spin tor que\nusing synthetic ferrimagnets\nChristian Klein\nSPSMS, UMR-E 9001 CEA / UJF-Grenoble 1, INAC, Grenoble, F-38 054, France,\nPhysikalisches Institut, Universit¨ at Karlsruhe (TH), Wo lfgang-Gaede Strasse 1, 76131 Karlsruhe, Germany\nCyril Petitjean and Xavier Waintal\nSPSMS, UMR-E 9001 CEA / UJF-Grenoble 1, INAC, Grenoble, F-38 054, France\n(Dated: November 27, 2018)\nWe discuss the current induced magnetization dynamics of sp in valves F 0|N|SyF where the free\nlayer is a synthetic ferrimagnet SyF made of two ferromagnet ic layers F 1and F 2coupled by RKKY\nexchange coupling. In the interesting situation where the m agnetic moment of the outer layer\nF2dominates the magnetization of the ferrimagnet, we find that the sign of the effective spin\ntorque exerted on the free middle layer F 1is controlled by the strength of the RKKY coupling:\nfor weak coupling one recovers the usual situation where spi n torque tends to, say, anti-align the\nmagnetization of F 1with respect to the pinned layer F 0. However for large coupling the situation\nis reversed and the spin torque tends to align F 1with respect to F 0. Careful numerical simulations\nin the intermediate coupling regime reveal that the competi tion between these two incompatible\nlimits leads generically to spin torque oscillator (STO) be havior. The STO is found in the absence\nof magnetic field, with very significant amplitude of oscilla tions and frequencies up to 50 GHz or\nhigher.\nPACS numbers: 72.25.Ba, 75.47.-m, 75.70.Cn, 85.75.-d\nSince the first successful experimental manipulation of\nmagnetic configurations by spin-polarized currents [1, 2],\nthe interest in spintronics devices entirely controlled by\nelectrical currents has rapidly increased [3–10, 12, 13].\nThe key concept behind theses experiments is the no-\ntion of spin transfer torque (STT) predicted by Slon-\nczewski[14] andBerger[15]: acurrentgetsspinpolarized\nbyafirst(pinned)magneticlayerandthentransferssome\nof its polarization to a second free layer, hence exerting\na torque on its magnetization. As a result the magneti-\nzation can switch to a different static configuration [2, 3]\nor even to a dynamical stationary regime where the mag-\nnetization of the free layers shows a sustained preces-\nsion [4–7]. This dynamical regime, known as the spin\ntorque oscillation (STO), allows to convert a dc voltage\ninto a microwaves signal at several GHz and has a real\npotential for applications (nanoscale microwave source\nor detector, high frequencies possibly above CMOS tech-\nnology, narrow band). There are still technological is-\nsues with STO based devices, however, such as very lim-\nited output power and the need of rather strong external\nfields. Indeed, the main mechanism for producing STO\nis based on a detailed tuning of the angular dependance\nof the spin torque [5] and requires the use of an exter-\nnal magnetic field. Recently, different mechanisms have\nemerged to enable STO behavior in the absence of exter-\nnal field: the first approach [9] relies on the precession\nof a non uniform magnetic structure, a magnetic vortex,\nthat can be excited at sub GHZ frequencies; a second\napproach [8, 16] uses a strongly asymmetric spin valve to\nincrease the anharmonicity of the angular dependance of\nthe spin torque, leading to the so-called ”wavy” behaviorof the STT [16–19]. Other approaches use an orthogonal\npolarizer [10] or a perpendicular free layer [11, 12].\nIn this letter we propose an entire new mechanism\nwhich naturally induces strong STO behavior, even in\nthe absence of magnetic field, using a synthetic ferrimag-\nnet (SyF) as the free layer of the spin valve. The SyF is\nmade of two ferromagnetic layers coupled through anti-\nferromagneticRKKYexchangeinteraction[20–22]. Here,\nwe predict that by simply tuning the effective strength\nof the RKKY exchange coupling (through the different\nthicknesses of the layers for instance), the signof the\neffective STT can be changed. At intermediate effective\ncoupling, the STT generically induces strong STO be-\nhavior irrespectively of the presence of applied magnetic\nfield.\nIn the following we first provide a simple physical pic-\nture that captures the essential features of our setup.\nOur main prediction uses very general conservation ar-\nguments and should therefore be quite robust and inde-\npendent of the details of the model. In a second step,\nwe provide more quantitative arguments to describe the\nstrong coupling regime. Last, we present a detailed nu-\nmerical study of the phase diagram of our model. We\nfind strong STO behavior at rather large frequencies in a\nwideportionofthephasediagramhintingthatSyFbased\nSTOs could be good candidates for application purposes.\nPhysical mechanism. We consider a nanopillar spin\nvalveF 0|N0|SyF made of apinned ferromagneticlayerF 0\nand a free synthetic ferrimagnetic layer SyF = F 1|N1|F2\n(where N 0and N 1are normal spacers), see Fig. 1a for a\ncartoon of the full stack. Upon injecting a current den-\nsityjthrough the nanopillar, a spin torque τττis exerted2\nFigure 1: Panel (a) is a cartoon of the magnetic multilayer\nstack, in which Miis the the magnetization of the layer F i.\nThe angle between M1andM0(M2) is respectively θ(α).\nPanels (b,c,d) show the stationary magnetization mz\n1along\nM0as a function of the current density jin A.cm−2, for a\nstack with d1= 2.nm,d2= 4d1and three different coupling\nstrength J; (b) weak coupling limit ( J= 0), (c) Intermediate\ncoupling limit ( J=−10−3J.m−2) (d) Strong coupling limit\n(J=−6.10−3J.m−2) Upper (Lower) inset : Sketch of the\ntorqueτττ1actionon M1(MSyF)fortheweak(strong)coupling\nregime.\non the magnetic moments of the layers [23]. This torque\nhas been extensively discussed in the literature and can\nbe strong enough to destabilize the initial magnetic con-\nfiguration. On the other hand, the magnetization M1\nandM2of F1and F 2are coupled by an oscillatory ex-\nchange (RKKY) interaction Jwhich is nothing but the\nspin torque present in the pillar at equilibrium [24, 25].\nThe coupling Jis reminiscent of Friedel oscillations and\ntypically behaves as J∼(cos2kFdN)/dNwherekFis\nthe Fermi momentum and dNthe width of the spacer\nN1.Jcan be tuned from negative (antiferromagnetic\ncoupling) to positive (ferromagnetic coupling), hence the\nname ”oscillatory”. Here we focus on negative values of\nJwhichstabilizean anti-alignedconfigurationof M1and\nM2. The goal of this letter is to discuss new phenomena\nthat arise from the competition between the equilibrium\nand the non equilibrium torques.\nIn order to gain a qualitative understanding of the un-\nderlying physics, let us discuss two extreme limits. We\nstart from a configuration where M1is aligned with M0whileM2is anti-aligned, and inject electrons from the\nright. Upon crossing the pinned layer F 0, the electrons\nget a polarization anti-parallel to M0. Denoting Jithe\nspincurrent(perunit surface)flowingin the normallayer\nNi, we have J0=/planckover2pi1pjm0/(2e) withm0=M0/|M0|and\npthe polarization of the current (0 ≤p≤1). Let us fo-\ncus on the weak coupling limit where J→0. When there\nis a finite angle θbetween M0andM1, the spin current\non the left J0and right J1of F1become different and\na torque τττ1=J0−J1is exerted on the magnetization\nof F1: conservation of angular momentum implies that\nwhatever spin lost by the conducting electrons is gained\nby magnetic degrees of freedom [14]. This conservation\nof angular momentum reads,\n∂\n∂t/bracketleftbiggd1M1\nγ1/bracketrightbigg\n=τττ1+... (1)\nwhered1is the width of F 1,γ1its gyromagnetic ratio\nand the ...indicate other contribution to the dynam-\nics (magnetic anisotropy, damping) to be discussed later\n(upper inset of Fig. 1). Ignoring (momentarily [26]) the\nrole of F 2, we arrive at ∂θ/∂t=γ1/planckover2pi1pjsinθ/(2ed1|M1|)\nand recover the usual phenomenology of spin torque: the\ntorque tends to stabilize the configuration where M1is\nanti-parallel to M0(or destabilize it if one reverses the\nsign of the current).\nLet us now discuss how this picture is modified when\none considers the opposite strong coupling J→ −∞\nlimit. Upon switching on J, the exchange energy (per\nunit surface of the pillar) EJ=−J(m1·m2) induces\na field like term in the RHS of Eq.(1) of the form\n+Jm2×m1. One needs also to consider the correspond-\ning equation for the dynamics of M2(see below Eq.(3)\nfor the full model) and one obtains a potentially rich cou-\npled dynamics for M1andM2. The situation simplifies\nin the limit where the exchange energy J→ −∞domi-\nnates: the relative configuration of M1andM2becomes\nfrozen in the anti-aligned position. Conservation of an-\ngular momentum now leads to a spin torque on the total\nmagnetic moment:\n∂\n∂t/bracketleftbiggd1M1\nγ1+d2M2\nγ2/bracketrightbigg\n=τττ1+... (2)\nwhere we have used the fact that τττ2=J1−J2vanishes\nwhenM1andM2are collinear. Note that the exchange\nfield, which conserves the total angular momentum, re-\ndistributes the torque τττ1onM1andM2so that they\nremain anti-aligned.\nEq.(2) allows two very different regimes: if\nd1|M1|/γ1> d2|M2|/γ2then the effective magnetiza-\ntion direction MSyFof the SyF layer points along M1\nand the torque is very similar to the weak coupling\nregime. This case has attracted some attention recently\nboth experimentally by Smith et al. [27] and theoreti-\ncally by Balaz et al. [28]. Here, we focus on the other3\nlimitd1|M1|/γ1< d2|M2|/γ2whereMSyFpoints in the\nopposite direction as M1(lower inset of Fig. 1). In\nthis limit, the torque tends to stabilize the configura-\ntion where MSyFis anti-aligned with M0hence where\nM0andM1are aligned: this is the opposite situation\nfrom the weak coupling J→0 limit.\nTo summarize, when going from J→0 toJ→ −∞,\nwithoutchangingthesignofthecurrent,thetorquetends\nto favor two different stationary situations. In the rest\nof this letter, we study this crossover in more details. In\nparticular, we find that the intermediate coupling regime\nreveals interesting, STO like, dynamical behavior.\nModel.We now turn to the modelisation of our sys-\ntem. The full stack we are considering has the form (see\nFig. 1a): AF |F0|N0|F1|N1|F2|N2, where AF is an antifer-\nromagnetic layer (typically IrMn) that pins the magne-\ntization of F 0. We set our reference spin axis ezparallel\ntoM0. Magnetization dynamics is described by two cou-\npled Landau Lifshiz Gilbert (LLG) equations that read\n(in SI units),\n∂m1\n∂t=γ1B1×m1+α1m1×∂m1\n∂t+γ1\n|M1|d1τττ1(3a)\n∂m2\n∂t=γ2B2×m2+α2m2×∂m2\n∂t+γ2\n|M2|d2τττ2,(3b)\nwheremi=Mi/|Mi|is the unit vector describing the\nmagnetization orientation of the layer Fiandαiis the\ndamping factor of the corresponding magnetic material.\nThe effective field Biseen by the magnetization in the\nlayer (i= (1,2)) is\nBi=Ki\nu\n|Mi|[mi·ez]ez+J\n|Mi|dim¯i,(4)\nwhere¯i= 3−icorresponds to the index of the other\nlayer. The effective field includes an uniaxial anisotropy\nKi\nufield and the RKKY exchange field. In the physical\nregimes studied below, RRKY coupling is the dominat-\ning energy so that our results are rather insensitive to\nthe presence of less important terms (for instance dipo-\nlar coupling between layers that we do not take into ac-\ncount). The spin torque τττiis calculated in the framework\nof CRMT (Continuous Random Matrix Theory) [16, 29].\nCRMT is a semiclassicalapproachthat generalizesValet-\nFert theory to non collinear situations. For discrete sys-\ntems, It is formally equivalent to the generalized circuit\ntheory [30] and is well adapted to the treatment of metal-\nlic magnetic multilayers.\nStrong coupling limit. When the RKKY coupling is\nlarge, one can derive an effective LLG equation for the\ndynamics of the SyF. It reads,\n∂m1\n∂t=γeffBeff×m1+αeffm1×∂m1\n∂t+γeff\n|M1|d1τττ1,(5)\nwhere the various effective parameters get renormalized\nas follows: γeff=γ1/(1−δ),αeff= (α1+δα2)/(1−δ),Keff\nu=K1\nu+K2\nud2/d1,Beff= (Keff\nu/|M1|)[m1·ez]ezand\nthe renormalization parameter δhas been defined as,\nδ=γ1|M2|d2\nγ2|M1|d1(6)\n(Eq.(5) is obtained by calculating ∂/∂t[m1+δm2] which\ncancels the RKKY contribution, and then setting m2=\n−m1andτττ2=0). From this equation we can clearly\nsee the inversion of the sign of the effective torque γeffτττ1\ndiscussed previously when δ >1.\nFigure 2: Upper: phase diagram of the system as a func-\ntion of the current density jand the thickness d1for a fixed\nδ= 4 and J=−5.10−3J.m−2. The various symbols indi-\ncate the observed stationary states in the corresponding re -\ngion (defined by different colors): ↑(↓) stands for mz\n1= 1\n(mz\n1=−1), while STOcorresponds to the presence a pre-\ncessional state. In the 2 STOregion, two different STO\nstates can be observed depending on the hysteresis history.\nLower: Resistance Ras function of time tfor a pillar surface\nof 1000 nm2andd1= 3nm. Upper (lower) plots correspond\nto initial conditions in the up (down) configuration. Left pa n-\nels:j= 2.8 107A.cm−2. Right panels: j= 5.8 107A.cm−2.\nThe amplitude of the oscillation with the up initial conditi on\nis small and invisible on the scale of the graphics\nNumerical simulations. Let us now go back to the full\nmodel Eq.(3) and study it numerically. We focus on a\nstack defined as IrMn 5|Py5|Cu10|Cod1|Ru0.3|Cod2|Cu10\nwhere the index indicate the widths of the layers in\nnm. We have used KCo\nu= 8.104J.m−3,MCo=\n1.42 106A.m−1and the transport parameters (spin re-\nsolved bulk resistivities, interface resistivities, spin flip\nlengths) have been taken from the database established\nin MSU and CNRS/Thales laboratory[31]. The coupling\nconstant Jcan be tuned by changing the Ruthickness\nwith typical values JRu≈ −5.10−3J.m−2[21]. The\ntorqueand resistanceshavebeen calculatedusing CRMT\nand the corresponding parametrization (as a function of\nthe angles θandα, see Fig.1a) incorporated into our nu-\nmerical LLG integrator.\nA first set of curve is presented in Fig. 1 b,c and d\nwhere the stationary magnetization mz\n1along the zaxis4\nis plotted against the current density jfor three values\nof the coupling constant: small (b), intermediate (c) and\nlarge coupling (d). For small coupling (b) , we recover\nthe usual behavior of current induced magnetic reversal\nin spin valves: at zero current the system has two sta-\nble configurations where m1is aligned ( mz\n1= 1) or anti\naligned ( mz\n1=−1) withm0. Upon injecting a strong\npositive current, one ”pushes” toward the anti-aligned\nconfiguration which becomes stable above a critical value\nof the current density j(0.5 107A.cm−2in this example).\nAt large coupling (d), this hysteretic curve is reversed, in\nagreement with the arguments developed above. A very\ninteresting regime is found at intermediate couplings (c)\nwhere one observes a rather large window where the sta-\ntionary value of mz\n1corresponds to a finite angle between\nm0andm1, i.e. to a dynamical state where a sustained\nprecession of m1is present (with frequencies around 15\nGHz in this example).\nThe coupling Jcan be varied by tuning the width dN\nof the normal (Ru) spacer. However, this is not very\neasy to control in practice, as the oscillatory character\nof the RKKY interaction make this variation non mono-\ntonic. Alternatively, one can explore the phase diagram\nof our stack by varying the thickness d1for a fixed ra-\ntioδ=d2/d1, hence varying the relative importance of\nbulk versus surface terms in the LLG equation. Fig. 2,\nshows the resulting phase diagram in the ( j,d1) plane for\nδ= 4. The different symbols indicate the possible sta-\ntionary states in the various regionsof the phase diagram\n(defined by the various colors). The presence of two dif-\nferent symbols (most regions)indicate that depending on\nthe hysteresis history, one or the other state is observed.\nIn particular, in the 2 STO region, two different STO\nstatescanbeobserveddepending, forinstance, ontheini-\ntial condition mz\n1(t= 0) =±1. These two different STO\nstates merge upon entering the 1 STO region. We have\ninvestigated other stacks with different material parame-\nters (not shown) and found very similar phase diagrams,\nincluding the presence of high frequency STO phases, as\nlong as one remains in the δ >1 regime. We observe two\nkind of STO behaviors, as evident from the Resistance\nversus time R(t) traces plotted in Fig. 2. When the am-\nplitude of the oscillating signal is small (lower left of Fig.\n2), we observe some beating behaviors with one well de-\nfined frequency and a second much smaller frequency less\nwell defined (corresponding to the precession of the two\nlayers respectively). When these two frequencies become\ncloser, the time dependent signal gets bigger and R(t)\nbecomes much more sinusoidal indicating phase locking\nbetween the precessional dynamics of the two magneti-\nzationsm1andm2.\nIn the last Fig.3, we focus on the region of the phase\ndiagram (lower right corner of Fig.2) where the ampli-\ntude of the STO signal is the highest. The upper curves\nshow the amplitude of the time dependent signal while\nthe lower curves show the corresponding frequency. Wefind that this large angle STO phase, which is stable in\nthe absence of magnetic field, leads to very high frequen-\ncies up to several tens of GHz while sustaining high pre-\ncession angles (of the order of π/6 ). The observed high\nfrequencies is tightly linked with the high value of the\nRRKY coupling which can be up to two orders of magni-\ntude larger than typical anisotropy or Zeeman energies.\nEveniftherelativeanglebetweenthetwomagnetizations\nof the synthetic ferrimagnet remain close to π, this gives\nrise to frequencies of several tens of GHz.\nToconclude, wehaveshownthatthe interplaybetween\nequilibriumandnonequilibriumtorqueinsyntheticferri-\nmagnets based spin valves can lead to interesting physics\nand potentially a new route to high frequency STOs.\nFigure 3: Upper plot: amplitude (in /permil) of the time depen-\ndent signal of the resistance as a function of d1for fixed\nj= 6.107A.cm−2. Lower plot: idem for the main oscillating\nfrequency of the resistance. Insets: corresponding colorp lots\nin the (j,d1) plane. δ= 4 and J=−5.10−3J.m−2.\nWe thank T. Rasing, W. Wulfhekel for very useful dis-\ncussions. This work was supported by , CEA NanoSim\nprogram, CEA Eurotalent and EC Contract Macalo.\n[1] M. Tsoi, A. G. M. Jansen, J. Bass, W.-C. Chiang,\nM. Seck, V. Tsoi, and P. Wyder, Phys. Rev. Lett. 80,\n4281 (1998).\n[2] E. Myers, D. Ralph, J. Katine, R. Louie, and\nR. Buhrman, Science 285, 867 (1999).\n[3] J. Katine, F. Albert, R. Buhrman, E. Myers, and\nD. Ralph, Phys. Rev. Lett. 84, 3149 (2000).\n[4] M. Tsoi, A. G. M. Jansen, J. Bass, W.-C. Chiang,\nV. Tsoi, and P. Wyder, Nature 406, 46 (2000).\n[5] S. I. Kiselev, J. C. Sankey, I. N. Krivorotov, N. C. Em-\nley, R. J. Schoelkopf, R. A. Buhrman, and D. C. Ralph,\nNature425, 380 (2003).5\n[6] W. Rippard, M. Pufall, S. Kaka, S. Russek, and T. Silva,\nPhys. Rev. Lett 92, 027201 (2004).\n[7] I. N. Krivorotov, N. C. Emley, J. C. Sankey, S. I. Kise-\nlev, D. C. Ralph, and R. A. Buhrman, Science 307, 228\n(2005).\n[8] O.Boulle, V.Cros, J.Grollier, L.G.Pereira, C.Deranlo t,\nF. Petroff, G. Faini, J. Barna, and A. Fert, Nature Phys.\n3, 492 (2007).\n[9] V. S. Pribiag, I. N. Krivorotov, G. D. Fuchs, P. M. Bra-\nganca, O. Ozatay, J. C. Sankey, D. C. Ralph, and R. A.\nBuhrman, Nature Phys. 3, 498 (2007).\n[10] D. Houssameddine, U. Ebels, B. Dela¨ eT, B. Rodmacq,\nI. Firastrau, F. Ponthenier, M. Brunet, C. Thirion, J. P.\nMichel, L. Prejbeanu-Buda, et al., Nature Mat. 6, 447\n(2007).\n[11] G. Consolo, L. Lopez-Diaz, L. Torrez, G. Finocchio, A.\nRomeo, and B. Azzerboni, Appl. Phys. Lett. 91, 162506\n(2007).\n[12] W. H. Rippard, A. M. Deac, M. R. Pufall, J. M. Shaw,\nM. W. Keller, S. E. Russek, and C. Serpico, Phys. Rev.\nB81, 014426 (2010).\n[13] M. Madami, S. Bonetti, G. Consolo, S. Tacchi, G. Car-\nlotti, G. Gubbiotti, F. B. Mancoff, M. A. Yar, and\nJ.˚Akerman, Nature Nanotechnology 6, 635 (2011). Na-\nture Nanotechnology 6, 635638 (2011)\n[14] J. C. Slonczewski, JMMM 62, L1 (1996).\n[15] L. Berger, Phys. Rev. B 54, 9353 (1996).\n[16] V. S. Rychkov, S. Borlenghi, H. Jaffres, A. Fert, and\nX. Waintal, Phys. Rev. Lett. 103, 066602 (2009).\n[17] J. Manschot, A. Brataas, and G. Bauer, Phys. Rev. B\n69, 092407 (2004).\n[18] J. Barna´ s, A. Fert, M. Gmitra, I. Weymann, and\nV. Dugaev, Phys. Rev. B 72, 024426 (2005).\n[19] M. Gmitra and J. Barnas, Phys. Rev. Lett. 96, 207205\n(2006).\n[20] S. S. P. Parkin, N. More, and K. P. Roche, Phys. Rev.\nLett.64, 2304 (1990).\n[21] S. Parkin, Phys. Rev. Lett. 67, 3598 (1991).\n[22] P. Bruno and C. Chappert, Phys. Rev. B 46, 261 (1992).\n[23]j >0correspondstoinjectinganelectrons from theright.\n[24] J. Slonczewski, Phys. Rev. B 39, 6995 (1989).\n[25] X. Waintal and P. W. Brouwer, Phys. Rev. B 65\n054407(2002).\n[26] In the case considered here, the effect of τττ2will only\nslightly increase the predicted effect.\n[27] N. Smith, S. Maat, M. J. Carey, and J. R. Childress,\nPhys. Rev. Lett. 101, 247205 (2008).\n[28] P. Bal´ aˇ z and J. Barna´ s, Phys. Rev. B 83, 104422 (2011).\n[29] S. Borlenghi, V. S. Rychkov, C. Petitjean, and X. Wain-\ntal, Phys. Rev. B 84, 035412 (2011).\n[30] G. E. W. Bauer, Y. Tserkovnyak, D. Huertas-Hernando,\nand A. Brataas, Phys. Rev. B 67, 094421 (2003).\n[31] H. Jaffres private communication." }, { "title": "2101.10240v1.Raman_Spectroscopy_and_Aging_of_the_Low_Loss_Ferrimagnet_Vanadium_Tetracyanoethylene.pdf", "content": "Raman Spectroscopy and Aging of the Low-Loss Ferrimagnet Vanadium\nTetracyanoethylene\nH. F. H. Cheung,1M. Chilcote,1H. Yusuf,2D. S. Cormode,2Y.\nShi,3M. E. Flatt\u0013 e,3E. Johnston-Halperin,2and G. D. Fuchs1, 4, ∗\n1School of Applied and Engineering Physics, Cornell University, Ithaca, NY 14853, USA\n2Department of Physics, The Ohio State University, Columbus, Ohio 43210, USA\n3Department of Physics and Astronomy, University of Iowa, Iowa City, Iowa 52242, USA\n4Kavli Institute at Cornell for Nanoscale Science, Ithaca, New York 14853, USA\nVanadium tetracyanoethylene (V[TCNE] x,x\u00192) is an organic-based ferrimagnet with a high\nmagnetic ordering temperature T C>600 K, low magnetic damping, and growth compatibility with\na wide variety of substrates. However, similar to other organic-based materials, it is sensitive to air.\nAlthough encapsulation of V[TCNE] xwith glass and epoxy extends the \flm lifetime from an hour\nto a few weeks, what is limiting its lifetime remains poorly understood. Here we characterize encap-\nsulated V[TCNE] x\flms using confocal microscopy, Raman spectroscopy, ferromagnetic resonance\nand SQUID magnetometry. We identify the relevant features in the Raman spectra in agreement\nwith ab initio theory, reproducing C = C ;C\u0011N vibrational modes. We correlate changes in the\ne\u000bective dynamic magnetization with changes in Raman intensity and in photoluminescence. Based\non changes in Raman spectra, we hypothesize possible structural changes and aging mechanisms in\nV[TCNE] x. These \fndings enable a local optical probe of V[TCNE] x\flm quality, which is invaluable\nin experiments where assessing \flm quality with local magnetic characterization is not possible.\nI. INTRODUCTION\nIn the \feld of coherent magnonics, where function is\nderived through the creation and manipulation of long-\nlived spin wave modes, yttrium iron garnet (YIG) is the\nprototype material due to its exceptionally low damping\nat room temperature. However, it is di\u000ecult to integrate\nwith other materials, requiring deposition on a lattice\nmatched substrate, e.g. gadolinium gallium garnet, and\na synthesis temperature above 800\u000eC [1{3]. In compari-\nson, vanadium tetracyanoethylene (V[TCNE] x,x\u00192) is\na low loss organic-based material with intrinsic damping\ncomparable to YIG, and it can be grown as a \flm with\nchemical vapor deposition at a mild temperature (50\u000eC)\nand on nearly arbitrary substrates [4, 5]. Furthermore,\nV[TCNE] xcan be patterned using electron-beam lithog-\nraphy lift-o\u000b techniques while retaining its low magnetic\nloss [6]. Hence it is an attractive alternative for magnon-\nics. In particular, it is a candidate for coupling spin wave\nmodes to isolated defect centers [7{10]. The ability to\nshape magnon modes with small volumes allows the re-\nalization of strong defect spin-magnon coupling [9].\nA practical challenge in working with V[TCNE] xis\nthat, like many organic-based and monolayer materials, it\nis air sensitive. The recent development of encapsulation\n∗Corresponding author: gdf9@cornell.edutechniques, similar to those used to protect organic light\nemitting diodes (OLEDs) [11{14], can extend its lifetime\nat room temperature and in ambient atmosphere from\nan hour to a few weeks [15]. However, the limitations in\nthe lifetime of V[TCNE] xand its associated aging mech-\nanisms are not well understood. For example, one aging\nmechanism is the chemical reaction with oxygen and wa-\nter, which turns a V[TCNE] x\flm from blue-green opaque\nto transparent [15]. This is not the only mechanism, as\nevidenced by aging observed in samples stored in an argon\nglove box and the slow change in the magnetic properties\nof encapsulated samples that remain opaque. Another\npiece of evidence is the slowing down of aging when the\nsample is stored at low temperature (-30\u000eC) in argon,\nsuggesting an internal change being the next dominant\naging mechanism.\nFirst we study accelerated aging through laser heating,\nwhich provides a clean condition where reactions with\noxygen and water are absent. Studying sample response\nto high intensity laser illumination is also of particular\nrelevance to proposed quantum interfaces between spin\nwaves and isolated defects, which are probed with a fo-\ncused laser beam. We observe a nonlinear dependence of\nlaser damage on optical power, which is consistent with a\nheating-based laser damage mechanism. Instead of being\nmerely a nuisance, this local laser damage opens up a new\navenue of V[TCNE] xpatterning, which is an alternative\nto electron-beam patterning [6]. We show preliminary\nresults on laser patterning of V[TCNE] xand discuss itsarXiv:2101.10240v1 [cond-mat.mtrl-sci] 25 Jan 2021resolution limit.\nTo better deploy V[TCNE] xoutside of a pristine envi-\nronment, e.g. inert atmosphere, we need a better under-\nstanding of its aging mechanisms. V[TCNE] xis amor-\nphous with a local structural order [16], making Ra-\nman spectroscopy an e\u000bective method for studying its\nstructural properties locally, at the micron scale. In\naddition to this optical probe, we concurrently mea-\nsure magnetic properties using ferromagnetic resonance\n(FMR) and superconducting quantum interference device\n(SQUID) magnetometry to correlate changes in chemical\nproperties with changes in magnetic properties.\nThe paper is structured as follows. We \frst describe\nsample fabrication and then explain the measurement\nprocedures for confocal microscopy, micro-focused Ra-\nman spectroscopy, FMR, and SQUID magnetometry.\nNext, we characterize V[TCNE] xusing the above meth-\nods. In particular, we show Raman spectra and explain\nthe associated vibrational modes and their relation to\nchemical structure. Next we study laser damage in detail,\nmonitoring how photoluminescence and Raman spectra\nchange with laser exposure. We show proof-of-concept\npatterning examples using laser damage to selectively re-\nmove magnetism. After measuring the optical and mag-\nnetic properties of pristine V[TCNE] x, we next study its\naging by monitoring how the above properties change in\ntime. In particular, we observe aged samples are more\nsusceptible to laser damage. We also identify Raman\npeaks observed in pristine samples which are absent in\naged or laser damaged samples.\nII. METHODS\nA. Sample growth\nV[TCNE] x\flms are grown by chemical vapor deposi-\ntion, similar to previous reports [4]. We examine 4 sam-\nples in the main text. The \frst sample (sample 1) is a\nuniform 400 nm V[TCNE] x\flm with 73 nm of aluminum\ndeposited with an evaporator, encapsulated with epoxy\nand a glass cover slip. V[TCNE] xthickness is estimated\nfrom growth time. The aluminum layer is designed to\nprevent the laser light from exciting epoxy, ensuring we\nare only probing V[TCNE] xRaman spectra and photo-\nluminescence.\nSample 2 is a 1.6 \u0016m thick sample. Sample 3 and 4 are\n400 nm thick samples grown in the same batch. All three\nsamples are encapsulated with glass and epoxy only.B.Ab initio calculations\nThe electronic structure and phonon modes of\nV[TCNE] xare calculated using the Vienna ab initio Sim-\nulation Package (VASP) (version 5.4.4) with a plane\nwave basis and projector-augmented-wave pseudopoten-\ntials [17{20]. These pseudopotentials use the generalized\ngradient approximation (GGA) of Perdew, Burke, and\nErnzerhof (PBE) [21]. A Hubbard constant U=4.19, de-\ntermined via a linear response method [22], was used\nin the phonon calculations. The hybrid functional\nHeyd{Scuseria{Ernzerhof (HSE) [23] with the standard\nrange separation parameter !=0.2 was also tested for this\nsystem and showed consistent results with the PBE+U\napproach.\nC. Optical and Magnetic Measurements\nOptical measurements including micro-focused Raman\nspectroscopy and photoluminescence are performed in a\nhomebuilt confocal microscope with a 532 nm continuous\nwave laser. Integrated light collected from the sample is\ndetected using a single photon counting module, whereas\nRaman spectra are recorded using a Princeton Instru-\nments spectrometer and low dark count camera.\nWe measure angle-resolved, \feld modulated FMR in\na homebuilt spectrometer. The setup consists of a mi-\ncrowave signal generator, an electromagnet, a pair of\nmodulation coils, a microwave diode detector and a lock\nin ampli\fer. Film magnetization is measured using a\nQuantum Design MPMS 3 SQUID magnetometer. Ap-\nplied magnetic \feld is in plane.\nIII. RAMAN SPECTROSCOPY\nWe measure Raman spectra to characterize V[TCNE] x\nchemical bonds and structure. To better understand\nthe Raman spectrum, we also perform density func-\ntional theory (DFT) calculations using VASP code. We\nshow the calculated V[TCNE] xstructure in \fgure 1(a),\nwhere the geometry agrees with previous DFT studies\n[24{26]. In \fgure 2, we plot the experimental Raman\nspectrum with DFT phonon density of states. Based\non comparison with DFT calculations, we assign 1300-\n1500 cm\u00001Raman peaks to C = C stretching modes, and\n2200 cm\u00001Raman peaks to C \u0011N stretching modes.\nExample vibrational modes are shown in \fgure 1(b),\ndisplaying modes with dominant C \u0011N (2352 cm\u00001) or\nC = C (1451 cm\u00001) bond stretching.\n2(b)\nV\nN\nC(a)FIG. 1. V[TCNE] xstructure and vibrational modes based on\ndensity functional theory (DFT) calculations. (a) V[TCNE] x\nstructure, showing 2 nonequivalent TCNE molecules. The\nbottom TCNE is bonded with 4 vanadium atoms ( \u00164 bonded)\nwhile the top TCNE is bonded with 2 vanadium atoms (trans-\n\u00162 bonded) (b) C \u0011N (2352 cm\u00001) and C = C (1451 cm\u00001)\nvibrational modes. Shown is the bottom ( \u00164 bonded) TCNE.\nHigh wavenumber vibrational modes ( >1000cm\u00001) are local-\nized on one or the other TCNE, with non-degenerate vibra-\ntional frequencies.\nNext, we explain the \fne Raman features and com-\npare them with previous studies. Fine features near 2202,\n2225 cm\u00001are in agreement with previous IR studies\n2194, 2214 cm\u00001[15, 27]. We observe 2121 cm\u00001in Ra-\nman spectroscopy, while previous IR spectroscopy only\nobserved a peak at 2155 cm\u00001[15], which could be ex-\nplained by di\u000berent IR and Raman activities of these\nvibrational modes. The Raman peaks at 1308, 1411,\n1530 cm\u00001are absent in IR spectrum, suggesting vana-\ndium atoms are more symmetrically bonded to the cen-\nter C = C bond, leading to low IR activity [27]. We at-\ntribute low wavenumber peaks at 336, 457, 543 cm\u00001to\nlow energy TCNE vibration modes and V \u0000N stretching\n/uni00000013/uni00000011/uni00000013/uni00000013/uni00000011/uni00000018/uni00000035/uni00000044/uni00000050/uni00000044/uni00000051/uni00000003/uni0000004c/uni00000051/uni00000057/uni00000048/uni00000051/uni00000056/uni0000004c/uni00000057/uni0000005c\n/uni0000000b/uni00000046/uni00000053/uni00000056/uni00000012/uni00000046/uni000000501/uni0000000c\n /uni0000000b/uni00000044/uni0000000c\n/uni00000013 /uni00000014/uni00000013/uni00000013/uni00000013 /uni00000015/uni00000013/uni00000013/uni00000013\n/uni0000003a/uni00000044/uni00000059/uni00000048/uni00000051/uni00000058/uni00000050/uni00000045/uni00000048/uni00000055/uni00000003/uni0000000b/uni00000046/uni000000501/uni0000000c\n/uni00000033/uni0000004b/uni00000052/uni00000051/uni00000052/uni00000051/uni00000003/uni00000027/uni00000032/uni00000036/uni0000000b/uni00000045/uni0000000cFIG. 2. V[TCNE] xRaman spectrum and density of states\n(DOS). (a) Experimental spectrum of pristine encapsulated\nV[TCNE] x. A broad baseline is subtracted from the raw Ra-\nman spectrum (see SI). (b) Ab initio V[TCNE] xphonon DOS.\nIn the plot, each mode is broadened as a Lorentzian with a\nFWHM of 20 cm\u00001.\nmodes.\nHere we have characterized Raman spectra of pristine\nencapsulated V[TCNE] xand assigned Raman peaks to\nparticular vibrational modes. Low wavenumber modes\nare particularly relevant for magnetism as they are in-\n\ruenced by V{N bonds. In subsequent sections, we mea-\nsure how these Raman peaks evolve under aging and laser\nheating induced damage, and show how Raman spectra\ncould be used to assess \flm quality.\nIV. PHOTOLUMINESCENCE AND LASER\nDAMAGE SUSCEPTIBILITY\nIn the following section, we study laser damage in a\npristine encapsulated V[TCNE] x\flm (sample 2). Using\nlaser-induced damage allows us to study aging in the ab-\nsence of oxygen and water. There is also independent\ninterest in studying V[TCNE] xunder focused laser in-\ntensity, in particular in relation to experiments coupling\nV[TCNE] xwith defect centers [9]. The goal here is to\ngain insights on the nature of aging and a quantitative\nmeasure of laser damage susceptibility. We note that the\nquantitative laser damage susceptibility is sample and\nencapsulation dependent and we focus on reproducible\ntrends across samples and relative changes.\nBecause the laser light is tightly focused, local temper-\nature can be large enough to cause degradation. Indeed,\n3(a)\n(c)(b)\n(d)FIG. 3. (a) Laser damage susceptibility \u001fPLshowing a nonlinear dependence on optical power. Also shown is a guide to the\neye showing an exponential dependence with power. Inset - photoluminescence time trace at an optical power 0.23 mW. (b)\nRaw Raman spectra taken at 0.10 mW after exposing the laser of di\u000berent powers for 4 minutes. (c)-(d) Baseline subtracted\nintegrated Raman intensity around 1400 cm\u00001. (d) Integrated Raman intensity around 2200 cm\u00001.\nphotoluminescence increases under continuous illumina-\ntion, which is accompanied by the V[TCNE] x\flm turning\ntransparent. Here we characterize laser damage suscepti-\nbility\u001fPLas the fractional rate of change in photolumi-\nnescence per unit optical power, namely\n\u001fPL=1\nPopt1\nPL(t= 0)d\ndtPL(t) (1)\nwherePLis the photoluminescence, Poptis the incident\nlaser power. We linear \ft the initial photoluminescence\nincrease (Fig. 3(a) inset) and extract \u001fPL.\nLaser damage susceptibility increases nonlinearly in op-\ntical power, with a threshold near 0.25 mW (Fig. 3(a)).\nA nonlinear laser damage susceptibility is consistent witha laser heating damage mechanism. While the temper-\nature rise is proportional to laser power, chemical reac-\ntion rates increase nonlinearly with temperature, causing\nmore rapid aging at a high optical power.\nNext we quantify how laser damage alters the Ra-\nman spectra. All measurements are done in a local area\n20\u0016m\u000220\u0016m, where the \flm is uniform. We suc-\ncessively illuminate di\u000berent spots with di\u000berent optical\npower for 4 minutes, and then measure Raman spectra\nwith a low laser power (93 \u0016W) for 4 minutes, with a\npeak laser intensity 5 \u0002104W=cm2. As the spot is ex-\nposed to higher laser power, the Raman intensity near\n2121 cm\u00001decreases with a corresponding increase in the\n2202, 2225 cm\u00001features (Fig. 3(b)). The side peaks at\n1308, 1530 cm\u00001of C=C bond increase in intensity and\n4linewidth. The lower wavenumber peaks <600 cm\u00001dis-\nappear. In addition, the total \ruorescence background\nincreases with increasing laser power. The quantitative\nchanges in Raman intensity are shown in \fgure 3(c,d).\nFocusing on qualitative features, the disappearance of\nlow wavenumber ( <600 cm\u00001) and the 2121 cm\u00001Ra-\nman peaks are clear, qualitative signatures of \flm aging.\nIn particular, low wavenumber vibrational modes have\nsigni\fcant V{N stretching components, therefore the dis-\nappearance of those peaks suggest a reduction in vana-\ndium bonding to TCNE groups. Likewise, the 2121 cm\u00001\nRaman peak corresponds to a C \u0011N vibrational modes,\nwhich is sensitive to vanadium nitrogen bonding and is\nrelevant for magnetic ordering in V[TCNE] x.\n10 μm\n(a)\n(b)\nFIG. 4. Laser patterning of a 400 nm thick V[TCNE] xsam-\nple (sample 4). (a) Photoluminescence map. High count rate\nregions are laser damaged, low count rate regions are the re-\nmaining undamaged material. (b) Grayscale optical micro-\ngraph of the same area. Laser damaged areas are more trans-\nparent and appear brighter in the image. This is patterned\nwith 1.7 mW laser power, corresponding to an optical inten-\nsity of 1:3\u0002106W=cm2.\nInstead of merely being a nuisance, laser damage could\nalso be used for patterning with micron-scale spatial ex-\ntent. We next explore using laser damage to pattern a\nV[TCNE] x\flm. If thermal degradation is nonlinear with\nlocal temperature rise, the attainable feature size could\nbe much smaller than the di\u000braction limit [28].\nWe show a proof-of-concept demonstration by laser\npatterning the authors' a\u000eliations on a V[TCNE] x\flm\n(Fig. 4(a)), where laser written area is damaged, produc-\ning a much higher photoluminescence rate. Based on the\nphotoluminescence map and the optical image, the laser\npatterned feature size is on the order of 1 \u0016m. Note the\nphotoluminescence map only measures optical properties,\nand the magnetic properties of laser patterned samples,\ne.g. magnetization pro\fle, supported spin wave modes,\nFIG. 5. Angle-resolved FMR measured at 3 GHz microwave\nfrequency. Fitted Heff= 68Oe. (Inset) Field sweep at 50\u000e.\nFitted FWHM linewidth is 1.1 Oe.\nwill be the topic of a future study.\nV. FERROMAGNETIC RESONANCE\nWe next measure the e\u000bective magnetization by angle-\nresolved FMR. The resonance frequency !with an ex-\nternal \feld much greater than the e\u000bective \feld of the\nmaterial (H\u001dHeff) is [4]\n!=\rq\n(H\u0000Heffcos2\u0012)(H\u0000Heffcos 2\u0012) (2)\nwhere\ris the gyromagnetic ratio, His the external mag-\nnetic \feld,\u0012is the angle of the external \feld with respect\nto the \flm normal, Heff= 4\u0019Meff= 4\u0019Ms\u0000Hk, which\nis a combination of the saturation magnetization and the\nanisotropy \feld. One possible origin of the anisotropy\n\feld is from strain due to di\u000berential thermal expansion\nof V[TCNE] xand the underlying substrate [29, 30]. In\n\fgure 5, we show a \feld-modulated FMR signal measured\nat 3 GHz. Fitting angle-resolved FMR, we extract an ef-\nfective \feld 4 \u0019Meff= 68 Oe. The FWHM linewidth is\n1.1 Oe, which is comparable to previously reported values\n[6].\nVI. AGING\nHaving characterized pristine encapsulated V[TCNE] x,\nwe next monitor its properties as it ages at room temper-\nature in ambient atmosphere. First, we note a visual\nchange in V[TCNE] x, beginning at the sample edge and\npropagating under the glass coverslip. This material is\ndark grey-black as deposited. As it ages, a transparency\n5Day 4\nDay 15 (a)\n100 μm\n(b)\n(c)FIG. 6. (a) Visual indication of V[TCNE] xaging, showing the\naging front advancing from day 4 to day 15. The V[TCNE] x\n\flm turns transparent and reveals the underlying re\rective\naluminum layer. (b) Raw Raman spectrum on day 1 (red),\nday 32 (blue) and day 89 (green) near sample center. The\noverall \ruorescence/Raman background \roor increases and a\nplateau arises near 1400 cm\u00001. Over long time (89 days), low\nwavenumber peaks (300-600 cm\u00001) and the peak near 2120\ncm\u00001vanishes. (c) Reduction in 4 \u0019Ms\u0000Hkand increase\nin laser damage susceptibility. Laser damage susceptibility\nmeasured at the center of the sample, using a laser power of\n100 - 200\u0016W. Fitted lines decay rate is 0.72 Oe/day and laser\ndamage susceptibility increases with a 1 =etime constant of 6.4\ndays.front propagates from the sample edges (Fig. 6(a)). We\nattribute this chemical change to reaction with di\u000bused\noxygen and water across the epoxy encapsulation barrier\nat sample edges.\nIn contrast, V[TCNE] xfar from the edges does not\nshow a strong color change with time, However, the lack\nof an apparent color change does not mean no changes\noccur. We next characterize sample photoluminescence\nand Raman spectra at the center of the sample and show\nhow they change in time. Raman spectra show an in-\ncrease in the \ruorescence background and an increase in\nthe Raman intensity near 1300 - 1500 cm\u00001as a function\nof time (Figure 6(b)). In addition, peaks near 1300, 1500\ncm\u00001increase in Raman intensity compared to the cen-\nter 1400 cm\u00001peak. Looking at an even longer timescale\n(89 days), the low wavenumber peaks (336, 457, 543\ncm\u00001) vanish, and the 2120 cm\u00001peak further dimin-\nishes. These features are qualitatively similar to those\nobserved in laser damaged samples, with the strongest\nchanges appearing in V{N and C \u0011N bonding. This sug-\ngests that the slow processes present in room temperature\naging are similar to the chemical reactions accelerated by\nlaser damage.\nA more drastic signature of aging is an increase in\nlaser damage susceptibility \u001fPL, which increases expo-\nnentially in time as the sample ages (Fig. 6(c)). Concur-\nrently, the e\u000bective magnetization Heff= 4\u0019Meff, as\nmeasured from angle-resolved FMR, decreases over time\n(Fig. 6(c)). This establishes a link between optical prop-\nerties and magnetic properties, indicating optical mea-\nsurements could be a local probe of V[TCNE] x\flm qual-\nity.\nOne limitation of the angle-resolved FMR is that it\nonly measures an intensive quantity Meff, which is a\nsum of shape anisotropy and other anisotropy, e.g. strain\ninduced anisotropy [30]. Moreover, it is not sensitive to\nthe total magnetic moment of the sample. To disentangle\nthese contributions and understand the total reduction of\nmagnetic material, we monitor aging of another 400 nm\nthick V[TCNE] x\flm (sample 3) with SQUID magnetom-\netry and FMR.\nHere we compare 3 quantities, total magnetic mo-\nmentmtotmeasured by SQUID magnetometry, e\u000bective\nmagnetization (4 \u0019Meff) and weighted total moment mw\nmeasured by angle-resolved FMR. Since mtotandmware\nextensive quantities while 4 \u0019Meffis an intensive quan-\ntity, we explain below in detail the measurement proce-\ndure for each quantity and how we compare them.\nWe compute saturation magnetization from SQUID\nmagnetometry data by normalizing the total moment to\nthe V[TCNE] x\flm volume. The volume has an uncer-\ntainty of up to 20%, which is dominated by the thickness\n6uncertainty. This study lacks a direct measure of the\nsample thickness for each sample, and there is growth-to-\ngrowth variation in the nominal deposition rate as well as\nvariation of growth rate across di\u000berent positions in the\ngrowth chamber. Nonetheless, we are most interested in\nhowMschanges in the aging process, which is less de-\npendent on initial sample volume. We also note that the\nvolume of ferrimagnetic V[TCNE] xis not constant as the\nsample ages due to the oxidation front that propagates\nfrom sample edge to center (see Fig. 6(a)), resulting in a\n33% reduction in opaque area over 32 days. In \fgure 7(b),\nwe plot the moment normalized to the initial V[TCNE] x\nvolume. We denote this quantity as 4 \u0019Msand we inter-\npret it as a scaled total ferrimagnetic moment.\nNext we discuss FMR measurements. We note the sam-\nple has a low loss fraction throughout the course of the ex-\nperiment, with a Gilbert damping varies from 1 :5\u000210\u00004\nto 2:5\u000210\u00004from day 1 to day 20 (Fig. 7(a)). We can\nunambiguously detect a resonance line shape and the typ-\nical uncertainty in 4 \u0019Meffis near 1 Oe. Because we\nmeasure FMR with \feld modulation, we are insensitive\nto high damping magnetic material, which will appear as\na broad background.\nThis weighted moment mwis computed based on\nFMR signal strength. With a microwave drive below\nsaturation and a low modulation \feld, the double in-\ntegrated signal is proportional to total magnetic mo-\nment [31, 32]. For an absorption derivative line shape\nL0\nabs(H) =a\u0001H3(H\u0000H0)\n(\u0001H2+4(H\u0000H0)2)2, this double integrated sig-\nnal is proportional to the product of signal amplitude\nand FWHM a\u0001H. We normalize a\u0001Hto microwave\npower and modulation amplitude (see SI) to compute\nmw, which is proportional to the magnetic moment. Be-\ncause the sample (2450 \u0016m) is several times wider than\nthe microwave waveguide (430 \u0016m), FMR sensitivity is\nnon-uniform across the sample. Hence we note mwis\na measure of weighted magnetic moment and is more\nsensitive to the sample portion closer to the microwave\nwaveguide. Moreover, the sample has been remounted\nmultiple times as we alternate between the SQUID and\nFMR setup, so the coupling between the V[TCNE] x\flm\nand the microwave waveguide varies, which contributes\nan uncertainty of nearly a factor of 2. With the above\nsubtleties pointed out, we use FMR intensity mwto esti-\nmate weighted magnetic moment with low damping.\nOver the course of the study (38 days), 4 \u0019Meffand\nmwreduces by nearly 2 orders of magnitude (Fig. 7(b)).\nIn comparison, 4 \u0019Meffonly changes from 95 \u00062 Oe to\n45\u00061 Oe. These \fndings suggest an aging process where\nthe magnetic portion shrinks while remaining low damp-\ning. This is consistent with an aging front propagatingfrom sample edge to center, where the center is relatively\npristine. These measurements suggest intrinsic aging due\nto internal chemical reaction at room temperature does\nnot increase damping signi\fcantly over 20 days.\nTo summarize the results of this section over several\nobservations, we \fnd a qualitative correlation between\nmagnetic properties and optical properties of V[TCNE] x.\nNamely, a decrease in 4 \u0019Meffcorrelates with the dis-\nappearance of Raman features (300-600 cm\u00001) involving\nV{N bonds and an increase in laser damage susceptibil-\nity. DC magnetometry, in combination with FMR, re-\nveals non-uniform aging resulting in a large decrease in\ntotal magnetic moment. The combination of optical and\nmagnetic measurements allow a comprehensive study of\nV[TCNE] xproperties, relating change in chemical prop-\nerties and magnetic properties.\nVII. CONCLUSION\nWe characterize pristine encapsulated V[TCNE] xthin\n\flms using confocal microscopy, micro-focused Raman\nspectroscopy, FMR and SQUID magnetometry. Through\ncomparison with ab initio calculations, we associate the\nexperimentally observed Raman peaks with particular\nC\u0011N;C = C stretching modes. We measure how the\nsample photoluminescence depends on laser power and\nobserve that laser damage susceptibility has a nonlinear\ndependence on laser power, which is consistent with a\nheating based laser damage mechanism. We further ex-\nplore laser damage as a means of patterning. Studying\nthe spatial pro\fle of laser damaged V[TCNE] xand their\nspin waves mode is a subject for future studies.\nWe identify changes in Raman features and laser dam-\nage susceptibility as sample ages. Low wavenumber (300-\n600 Cm\u00001) Raman features associated with V{N bonds\nvanish as sample ages, suggesting changes in bonding be-\ntween vanadium and TCNE. These \fndings show that\noptical measurement is a local probe of V[TCNE] x\flm\nmagnetic quality and could assess magnetic microstruc-\nture quality.\nThe existence of a narrow FMR response in encapsu-\nlated V[TCNE] xover 20 days under ambient conditions\nsuggests that intrinsic aging (e.g. not oxidation) does not\nincrease damping over this time interval. This is promis-\ning for coherent magnonics using V[TCNE] xmicrostruc-\ntures that are positioned far from encapsulation edges.\nFor quantum applications that use V[TCNE] xat cryo-\ngenic temperatures, one expects that the damping prop-\nerties will be preserved over an even longer timescale.\n7(a)\n(b)FIG. 7. Magnetic properties of an encapsulated V[TCNE] x\n\flm as it ages under ambient conditions. (a) The Gilbert\ndamping parameter \u000bextracted from FMR measurements in\nthe range 2-11 GHz with an out of plane DC magnetic \feld.\n(Inset) Gilbert damping \ft on day 1. \u000b= (1:8\u00060:1)\u000210\u00004\n(b) Magnetic measurements vs. sample aging time. SQUID\nmagnetometry measurements and FMR intensity consistently\nshow a large decrease in magnetic moment. In comparison,\n4\u0019Meff= 4\u0019Ms\u0000Hkshows a much smaller change over the\nsame period of time.\nVIII. ACKNOWLEDGEMENTS\nOptical Raman experiments, optical writing experi-\nments, waveguide FMR studies of aging, \frst principles\ntheory, and sample growth of aging study samples were\nsupported by the US Department of Energy O\u000ece of\nScience, O\u000ece of Basic Energy Sciences under Award\nDE-SC0019250. Control sample growth and cavity FMR\ncharacterization, as well as growth of 2{4 year old sam-\nples were supported by NSF DMR-1808704 and DMR-\n1507775. We acknowledge use of the facilities of the Cor-nell Center for Materials Research, which is supported\nthrough the NSF MRSEC program (DMR-1719875), and\nthe Cornell NanoScale Facility, which is a member of\nthe National Nanotechnology Coordinated Infrastructure\nand supported the NSF (NNCI-2025233). We acknowl-\nedge helpful discussions with Brendan McCullian. We\nthank Hong Tang and Na Zhu for providing 2{4 year old\nV[TCNE] xreference samples.\nREFERENCES\n[1] Christoph Hauser, Tim Richter, Nico Homonnay, Chris-\ntian Eisenschmidt, Mohammad Qaid, Hakan Deniz, Di-\netrich Hesse, Maciej Sawicki, Stefan G. Ebbinghaus, and\nGeorg Schmidt. Yttrium iron garnet thin \flms with very\nlow damping obtained by recrystallization of amorphous\nmaterial. Scienti\fc Reports , 6(1):20827, Feb 2016.\n[2] Yiyan Sun, Young-Yeal Song, Houchen Chang, Michael\nKabatek, Michael Jantz, William Schneider, Mingzhong\nWu, Helmut Schultheiss, and Axel Ho\u000bmann. Growth\nand ferromagnetic resonance properties of nanometer-\nthick yttrium iron garnet \flms. Applied Physics Letters ,\n101(15):152405, 2012.\n[3] Tao Liu, Houchen Chang, Vincent Vlaminck, Yiyan Sun,\nMichael Kabatek, Axel Ho\u000bmann, Longjiang Deng, and\nMingzhong Wu. Ferromagnetic resonance of sputtered\nyttrium iron garnet nanometer \flms. Journal of Applied\nPhysics , 115(17):17A501, 2014.\n[4] H. Yu, M. Harberts, R. Adur, Y. Lu, P. Chris Ham-\nmel, E. Johnston-Halperin, and A. J. Epstein. Ultra-\nnarrow ferromagnetic resonance in organic-based thin\n\flms grown via low temperature chemical vapor depo-\nsition. Applied Physics Letters , 105(1):012407, 2014.\n[5] Na Zhu, Xufeng Zhang, I. H. Froning, Michael E. Flatt\u0013 e,\nE. Johnston-Halperin, and Hong X. Tang. Low loss spin\nwave resonances in organic-based ferrimagnet vanadium\ntetracyanoethylene thin \flms. Applied Physics Letters ,\n109(8):082402, 2016.\n[6] Andrew Franson, Na Zhu, Seth Kurfman, Michael\nChilcote, Denis R. Candido, Kristen S. Buchanan,\nMichael E. Flatt\u0013 e, Hong X. Tang, and Ezekiel\nJohnston-Halperin. Low-damping ferromagnetic res-\nonance in electron-beam patterned, high-q vanadium\ntetracyanoethylene magnon cavities. APL Materials ,\n7(12):121113, 2019.\n[7] Toeno van der Sar, Francesco Casola, Ronald Walsworth,\nand Amir Yacoby. Nanometre-scale probing of spin waves\nusing single electron spins. Nature Communications ,\n6(1):7886, Aug 2015.\n[8] Paolo Andrich, Charles F. de las Casas, Xiaoying Liu,\nHope L. Bretscher, Jonson R. Berman, F. Joseph Here-\nmans, Paul F. Nealey, and David D. Awschalom. Long-\nrange spin wave mediated control of defect qubits in nan-\nodiamonds. npj Quantum Information , 3(1):28, Jul 2017.\n8[9] Denis Ricardo Candido, Greg D Fuchs, Ezekiel Johnston-\nHalperin, and Michael Flatt\u0013 e. Predicted strong coupling\nof solid-state spins via a single magnon mode. Materials\nfor Quantum Technology , 2020.\n[10] Iacopo Bertelli, Joris J. Carmiggelt, Tao Yu, Brecht G.\nSimon, Coosje C. Pothoven, Gerrit E. W. Bauer,\nYaroslav M. Blanter, Jan Aarts, and Toeno van der Sar.\nMagnetic resonance imaging of spin-wave transport and\ninterference in a magnetic insulator. Science Advances ,\n6(46), 2020.\n[11] Myeon-Cheon Choi, Youngkyoo Kim, and Chang-Sik Ha.\nPolymers for \rexible displays: From material selection\nto device applications. Progress in Polymer Science ,\n33(6):581 { 630, 2008.\n[12] Anna B. Chwang, Mark A. Rothman, Sokhanno Y. Mao,\nRichard H. Hewitt, Michael S. Weaver, Je\u000b A. Silvernail,\nKamala Rajan, Michael Hack, Julie J. Brown, Xi Chu,\nLorenza Moro, Todd Krajewski, and Nicole Rutherford.\nThin \flm encapsulated \rexible organic electrolumines-\ncent displays. Applied Physics Letters , 83(3):413{415,\n2003.\n[13] S. P. Subbarao, M. E. Bahlke, and I. Kymissis. Labora-\ntory thin-\flm encapsulation of air-sensitive organic semi-\nconductor devices. IEEE Transactions on Electron De-\nvices, 57(1):153{156, 2010.\n[14] Rakhi Grover, Ritu Srivastava, Omwati Rana, D. S.\nMehta, and M. N. Kamalasanan. New organic thin-\flm\nencapsulation for organic light emitting diodes. Journal\nof Encapsulation and Adsorption Sciences , 01(02):23{28,\n2011.\n[15] I. H. Froning, M. Harberts, Y. Lu, H. Yu, A. J. Epstein,\nand E. Johnston-Halperin. Thin-\flm encapsulation of the\nair-sensitive organic-based ferrimagnet vanadium tetra-\ncyanoethylene. Applied Physics Letters , 106(12):122403,\n2015.\n[16] D. Haskel, Z. Islam, J. Lang, C. Kmety, G. Srajer, K. I.\nPokhodnya, A. J. Epstein, and Joel S. Miller. Local struc-\ntural order in the disordered vanadium tetracyanoethy-\nlene room-temperature molecule-based magnet. Phys.\nRev. B , 70:054422, Aug 2004.\n[17] G. Kresse and J. Hafner. Ab initio molecular dynamics\nfor liquid metals. Phys. Rev. B , 47:558{561, Jan 1993.\n[18] G. Kresse and J. Hafner. Ab initio molecular-dynamics\nsimulation of the liquid-metal{amorphous-semiconductor\ntransition in germanium. Phys. Rev. B , 49:14251{14269,\nMay 1994.\n[19] G. Kresse and J. Furthm uller. E\u000eciency of ab-initio total\nenergy calculations for metals and semiconductors using\na plane-wave basis set. Computational Materials Science ,\n6(1):15 { 50, 1996.\n[20] G. Kresse and J. Furthm uller. E\u000ecient iterative schemes\nfor ab initio total-energy calculations using a plane-wave\nbasis set. Phys. Rev. B , 54:11169{11186, Oct 1996.\n[21] John P. Perdew, Kieron Burke, and Matthias Ernzerhof.\nGeneralized gradient approximation made simple. Phys.\nRev. Lett. , 77:3865{3868, Oct 1996.[22] Matteo Cococcioni and Stefano de Gironcoli. Linear re-\nsponse approach to the calculation of the e\u000bective inter-\naction parameters in the LDA + U method. Phys. Rev.\nB, 71:035105, Jan 2005.\n[23] Jochen Heyd, Gustavo E. Scuseria, and Matthias Ernzer-\nhof. Hybrid functionals based on a screened coulomb po-\ntential. The Journal of Chemical Physics , 118(18):8207{\n8215, 2003.\n[24] Giulia C. De Fusco, Leonardo Pisani, Barbara Monta-\nnari, and Nicholas M. Harrison. Density functional study\nof the magnetic coupling in V(TCNE)2.Phys. Rev. B ,\n79:085201, Feb 2009.\n[25] Fanica Cimpoesu, Bogdan Frecus, Corneliu I. Oprea, Pe-\ntre Panait, and Mihai A. G^ \u0010rt \u0018u. Disorder, exchange and\nmagnetic anisotropy in the room-temperature molecular\nmagnet v[tcne]x { a theoretical study. Computational\nMaterials Science , 91:320 { 328, 2014.\n[26] Bogdan Frecus, Corneliu I. Oprea, Petre Panait, Mar-\nilena Ferbinteanu, Fanica Cimpoesu, and Mihai A.\nG^ \u0010rT \u0018 u. Ab initio study of exchange coupling for the con-\nsistent understanding of the magnetic ordering at room\ntemperature in v[tcne]x. Theoretical Chemistry Accounts ,\n133(5):1470, Mar 2014.\n[27] K. I. Pokhodnya, A. J. Epstein, and J. S. Miller. Thin-\n\flm v[tcne]x magnets. Advanced Materials , 12(6):410{\n413, 2000.\n[28] Quang Cong Tong, Dam Thuy Trang Nguyen,\nMinh Thanh Do, Mai Hoang Luong, Bernard Jour-\nnet, Isabelle Ledoux-Rak, and Ngoc Diep Lai. Direct\nlaser writing of polymeric nanostructures via optically\ninduced local thermal e\u000bect. Applied Physics Letters ,\n108(18):183104, 2016.\n[29] Michael Chilcote, Megan Harberts, Bodo Fuhrmann,\nKatrin Lehmann, Yu Lu, Andrew Franson, Howard\nYu, Na Zhu, Hong Tang, Georg Schmidt, and Ezekiel\nJohnston-Halperin. Spin-wave con\fnement and coupling\nin organic-based magnetic nanostructures. APL Materi-\nals, 7(11):111108, 2019.\n[30] Huma Yusuf, Michael Chilcote, Denis R. Candido,\nSeth W. Kurfman, Donley S. Cormode, Yu Lu,\nMichael E. Flatt\u0013 e, and Ezekiel Johnston-Halperin. Tem-\nperature dependent anisotropy and linewidth in the high-\nq ferrimagnet v(tcne)x, 2020.\n[31] Charles P. Poole. Electron spin resonance: a compre-\nhensive treatise on experimental techniques . Interscience\nPublishers, New York, 1967.\n[32] Gareth R. Eaton, Sandra S. Eaton, David P. Barr,\nand Ralph Thomas Weber, editors. Quantitative EPR .\nSpringer, Wien ; New York, 2010. OCLC: ocn310400817.\n[33] Hakan Urey. Spot size, depth-of-focus, and di\u000braction\nring intensity formulas for truncated gaussian beams.\nAppl. Opt. , 43(3):620{625, Jan 2004.\n[34] Princeton Instruments. IntelliCal FAQ , 9 2012.\n[35] Sung-June Baek, Aaron Park, Young-Jin Ahn, and Jae-\nbum Choo. Baseline correction using asymmetrically\nreweighted penalized least squares smoothing. Analyst ,\n9140:250{257, 2015.\n[36] C. Le Losq. Rampy.\n[37] G Kresse and J Hafner. Norm-conserving and ultrasoft\npseudopotentials for \frst-row and transition elements.\nJournal of Physics: Condensed Matter , 6(40):8245{8257,\noct 1994.\n[38] G. Kresse and D. Joubert. From ultrasoft pseudopoten-\ntials to the projector augmented-wave method. Phys.\nRev. B , 59:1758{1775, Jan 1999.\n[39] Sangita S. Kalarickal, Pavol Krivosik, Mingzhong Wu,\nCarl E. Patton, Michael L. Schneider, Pavel Kabos, T. J.\nSilva, and John P. Nibarger. Ferromagnetic resonance\nlinewidth in metallic thin \flms: Comparison of measure-\nment methods. Journal of Applied Physics , 99(9):093909,\n2006.\n[40] E. R. J. Edwards, A. B. Kos, M. Weiler, and T. J. Silva.\nA microwave interferometer of the michelson-type to im-\nprove the dynamic range of broadband ferromagnetic res-\nonance measurements. IEEE Magnetics Letters , 8:1{4,\n2017.\n[41] TJ Silva, HT Nembach, JM Shaw, B Doyle, K Oguz,\nK O'brien, and M Doczy. Characterization of magnetic\nnanostructures for spin-torque memory applications with\nmacro-and micro-scale ferromagnetic resonance. In Char-\nacterization and Metrology for Nanoelectronics . Singa-\npore: Pan Stanford Publishing, 2016.\n[42] Hoosung Lee. Rapid measurement of thermal conductiv-\nity of polymer \flms. Review of Scienti\fc Instruments ,\n53(6):884{887, 1982.\n[43] Shouhang Li, Xiaoxiang Yu, Hua Bao, and Nuo Yang.\nHigh thermal conductivity of bulk epoxy resin by bottom-\nup parallel-linking and strain: A molecular dynam-\nics study. The Journal of Physical Chemistry C ,\n122(24):13140{13147, 2018.\n[44] Quantum Design. Correcting for the Absolute Field Error\nusing the Pd Standard , 3 2018. Rev. A0.IX. SUPPLEMENTAL INFORMATION\nA. Confocal measurement\nV[TCNE] xsamples are measured in a homebuilt con-\nfocal microscope with a 532 nm continuous wave laser \fl-\ntered with a bandpass \flter (Iridian 532 BPF, ZX000163).\nIt is focused with a 50x, 0.7 NA objective (Olympus\nLCPLFLN50xLCD) on the sample. Re\rection and \ru-\norescence from the sample is collected and then \fl-\ntered by a long pass \flter (Iridian 532 LPF nano cuto\u000b,\nZX000850). This light is split between a silicon avalanche\nphotodiode (Excelitas SPCM-AQRH-13-FC) and a spec-\ntrometer (Princeton Acton SP-2500, focal length 500 nm\nwith a 300 g/mm grating) with a low dark count camera\n(PyLoN 100 BR).\nAn incident laser beam of 5 mm 1 =e2diameter illumi-\nnates the objective back aperture (5 mm diameter). At\nthis truncation ratio, the di\u000braction limited 1 =e2beam\ndiameter is 700 nm [33]. At a laser power 500 \u0016W, peak\nlaser intensity is 2 :6\u0002105W=cm2.\nB. Raman spectroscopy\nThe spectrometer is calibrated with IntelliCal wave-\nlength source and intensity source [34]. Absolute Ra-\nman spectra wavelength accuracy is limited by excita-\ntion laser wavelength accuracy, which limits the Raman\nspectroscopy accuracy to 4-8 cm\u00001.\nFor samples that are more susceptible to laser damage,\nwe reduce laser power and raster scan across \u0018100\u0016m2\nto reduce accumulated laser damage.\nWe remove cosmic rays from Raman spectra by re-\nmoving data points that have counts higher than its\nneighbor above a threshold. The Raman baseline is \ftt\nwith asymmetrically reweighted penalized least squares\nsmoothing [35] with the implementation in the Python\nmodule Rampy [36]. Processed Raman spectra are \ftted\nwith Lorentzian peaks.\nC.Ab initio calculations\nThe electronic structure and phonon modes of\nV[TCNE] xare calculated using the Vienna ab initio Sim-\nulation Package (VASP) (version 5.4.4) , which uses a\nplane-wave basis and pseudopotentials [17{20]. The pseu-\ndopotentials we used are default options from VASP's of-\n\fcial PAW potential set, with 5 valance electrons per V,\n104 valance electrons per C, and 5 valance electrons per N\natom [37, 38].\nThese pseudopotentials use the generalized gradient\napproximation (GGA) of Perdew, Burke, and Ernzerhof\n(PBE) [21]. A PBE+U approach with U=4.19 was used\nin the phonon calculation; with U determined via a linear\nresponse method [22]. This approach is chosen as simple\nGGA calculations fail to capture the d-orbital behavior\nof the V atom in our electronic structure calculations.\nHybrid functional Heyd{Scuseria{Ernzerhof (HSE) [23]\nwith standard range separation parameter !=0.2 was also\ntested for this system and showed consistent results with\nthis PBE+U approach. For the rest of the calculation,\nwe used 400 eV for the energy cuto\u000b, and a \u0000-centered\n2\u00022\u00022 k-mesh sampling.\nThe geometry of V[TCNE] xis consistent with the work\nof De Fusco et al. [24]. The unit cell has a triclinic\nstructure and consists of 1 V atom, 12 C atoms and 8 N\natoms. The structure used in the phonon calculation is\nour own relaxed system under the same set of parameters.\nD. Ferromagnetic resonance\nSample 1 is placed on a 50 \u0016m inner diameter loop an-\ntenna. A microwave signal is applied to the antenna with\na signal generator (Anristu MG3692C). The re\rected sig-\nnal is routed by a circulator (Narda-MITEQ Model 4923,\n2 - 4 GHz) to a diode detector (Herotek DHM185AB).\nWe sweep the magnetic \feld at a \fxed microwave fre-\nquency (3 GHz typical) and modulate an external \feld\nat a modulation frequency of 587 Hz. Detected power is\nfed into a lock in ampli\fer (Signal Recovery 7265) and\ndetected at the modulation frequency.\nSample 3 is placed on a microstrip test board (South-\nwest Microwave B4003-8M-50). We launch microwave\nsignal at one end and do lock in detection of the trans-\nmitted microwave power.\nAs the sample ages, we increase modulation \feld am-\nplitude and microwave power to maintain a large enough\nsignal to noise.\nE. FMR data analysis\nThe spectra are \ftted with sums of Lorentzian deriva-\ntives, each having an absorptive and a dispersive part\n[39].L0\nabs(H) =a\u0001H3(H\u0000H0)\n(\u0001H2+ 4(H\u0000H0)2)2(3)\nL0\ndisp(H) =d\u0001H2(\u0001H2\u00004(H\u0000H0)2)\n(\u0001H2+ 4(H\u0000H0)2)2(4)\nwhere \u0001His the full-width at half-maximum of the\nLorentzian.\nIn angle-resolved FMR, we \ft with the following func-\ntion, using \r;H eff;\u00120set as free parameters.\n!=\rq\n(H\u0000Heffcos2(\u0012\u0000\u00120))(H\u0000Heffcos 2(\u0012\u0000\u00120))\n(5)\nwhere\ris the gyromagnetic ratio, Heff= 4\u0019Meffis\nthe e\u000bective magnetization, and \u00120is a constant o\u000bset\nbetween the real \flm normal with respect to the nominal\n\flm normal.\nNext we explain the procedure of extracting the total\nmagnetic moment from the FMR signal strength. The\nchange in waveguide transmission parameter on magnetic\nresonance is [40, 41]\n\u0001S21(Hres)/\r\u00160Msldm\n8Z0\u000bw(6)\nwherelis the \flm length along the transmission line, dm\nis the \flm thickness, Z0is the characteristic impedance\nof the transmission line, and \u000bis the Gilbert damping.\nAt a \fxed resonance \feld, the absorption curve FWHM is\n\u0001Hfwhm = 2\u000bHres. Hence the integrated area under the\nabsorption curve with respect to external H\feld is pro-\nportional to a weighted magnetic moment, independent\nof damping.\nFrom the absorption derivative term L0\nabs(H) =\na\u0001H3(H\u0000H0)\n(\u0001H2+4(H\u0000H0)2)2, we extract an equivalent area aH0as\nan estimate of the total magnetic moment.\nWe adjust for microwave power by normalizing the de-\ntected signal to microwave power ( P0) and the power-to-\nvoltage conversion factor ( K(V/mW)) of the diode de-\ntector. Namely, the normalized signal is\nVsig\nP0K(7)\nIn modulation detection, when modulation \feld is\nmuch smaller than linewidth, the lock in detected signal\nis proportional to a derivative of a Lorentzian absorp-\ntion. When modulation \feld is comparable to linewidth,\ndetected signal is no longer a simple derivative and we\n11correct for that by modeling the line shape.\nNote the FMR line shape has both an absorptive\nderivative and a dispersive derivative. Here we only use\nthe absorptive derivative to extract the total magnetic\nmoment.\nF. Laser heating\nIn this section, we estimate local temperature rise due\nto laser heating of a V[TCNE] x\flm grown on a glass\nsubstrate. We approximate the glass substrate thermal\nconductivity to be \u00141= 1:4Wm\u00001K\u00001. We approx-\nimate V[TCNE] xand epoxy to have a similar thermal\nconductivity \u00142= 0:3Wm\u00001K\u00001[42, 43].\nBecause the laser spot size (350 nm) is much smaller\nthan the epoxy thickness (5-10 \u0016m), we model both the\nsubstrate and V[TCNE] xplus epoxy as semi-in\fnite. The\ntemperature rise pro\fle for a point heat source at the\ninterface is\nT(r) =1\n4\u0019\u0016\u0014_Q\nr(8)\nwhere \u0016\u0014= (\u00141+\u00142)=2, assuming temperature is \fxed at\n0 at in\fnity.\nWe model the actual heat source as a surface Gaussian\nheat source\n2P0\n\u0019w2exp\u0012\n\u00002r2\nw2\u0013\n(9)\nwherewis the 1=e2beam radius.\nThe maximum temperature rise occurs at the interface\nalong the beam center,\nTmax=1\n2p\n2\u0019\u0016\u0014P0\nw(10)\nAt 0.7 NA, 1 =e2beam radius w\u0019350 nm. 100 \u0016W of\npower causes a local temperature rise of 70 K.\nA more realistic estimate takes into account of the \fnite\nabsorption depth of V[TCNE] x, which is of the order 100-\n400 nm. As this is comparable to the beam size, the peak\ntemperature rise is of the same order of magnitude.\nG. Long term aging\nVisual inspection provides preliminary information on\nV[TCNE] x\flm quality, but it is neither a quantitative\nnor a de\fnitive measure. While discolored V[TCNE] xis almost certainly aged, even fully opaque \flm can be\nmagnetically inactive[15].\nTo better understand long term degradation, we ex-\namine 2 other V[TCNE] xsamples. One is a 4 year old\nuniform \flm, the other is a 2 year old, patterned 100\n\u0016mwide bars. Both of them are nominal 1 \u0016m thick,\nestimated from growth time. Both of them are visually\nopaque and yet are magnetically inactive.\nThe samples show qualitatively similar peaks (Fig. 8),\nwith 3 peaks near 1300-1500 cm\u00001and a single peak\nnear 2200 cm\u00001. Under strong laser illumination, 3 peaks\n1300-1500 cm\u00001merge into 2 peaks, and become brighter\nthan the 2200 cm\u00001peak. Crucially, the spectra di\u000ber\nsigni\fcantly from aged V[TCNE] x. Degraded samples\ndon't have low wavenumber peaks (300-600 cm\u00001) and\nthe small 2120 cm\u00001peak as seen in pristine samples,\nfurther supporting a link between these \fne features with\ngood magnetic properties.\nH. SQUID magnetometry\nWe measure a 400 nm thick V[TCNE] x\flm (sample 3)\nin a Quantum Design MPMS 3 SQUID magnetometer in\nthe vibrating sample magnetometer (VSM) mode. The\nsample is mounted to a quartz holder with a small dab\nof GE varnish. We measure moment vs. \feld at 300 K\nin a four-quadrant sweep, from {20,000 Oe to +20,000\nOe back to {20,000 Oe. We use a palladium reference to\ncorrect for the absolute \feld error [44]. The applied \feld\nis in the plane of the V[TCNE] x\flm.\nWe normalize the measured magnetic moment to the\ninitial \flm volume, which is estimated from the opaque\n\flm area (7.27 mm2) and the nominal \flm thickness 400\nnm.\nThe magnetization vs. \feld data has a large diamag-\nnetic background (Fig. 9(c)), which we subtracted away\nby \ftting to a linear background at \felds jHj>2000\nOe. We also measure the magnetic response GE varnish,\nwhich has a small diamagnetic and ferromagnetic compo-\nnent (Fig. 9(a,b)). The magnetic moment of GE varnish\nis normalized to the same initial V[TCNE] xvolume. Note\nthat this is an equivalent V[TCNE] xmagnetization error\ndue to GE varnish background but not the physical GE\nvarnish magnetization. As the amount of GE varnish ap-\nplied in di\u000berent runs varies, one cannot simply subtract\na \fxed background from the data. We note GE varnish\nhas a gradual moment vs. \feld dependence over {1000\nOe to +1000 Oe (Fig. 9(b)). Because the applied \feld is\nin plane, V[TCNE] xshould be close to fully magnetized\nat 200 Oe. Therefore, we estimate V[TCNE] x4\u0019Msto\nbe the magnetization at 200 Oe, where background mag-\n12(a)\n(d)\n(g)(b)\n(e)\n(h)(c)\n(f)\n(i)\nFIG. 8. Long term study of encapsulated V[TCNE] xRaman spectra. The Raman spectra are baseline subtracted. We probe\nthe center opaque areas away from encapsulation edges in all 3 samples. (a) 4 year old, 1 \u0016m thick uniform \flm. (b) 2 year old,\n1\u0016m thick, patterned 100 \u0016m wide bar sample. (c) (Sample 1 in the main text) 89 days old, 400 nm thick \flm with aluminum\nencapsulation. Note that the V[TCNE] x\flm has been remounted on the antenna and their relative position di\u000ber from that\nin the main text. (d)-(f) The laser is raster scanned across near a 50 \u0016m2area to reduce laser damage. Laser power is (d) 190\n\u0016W, (e) 190 \u0016W, (f) 51\u0016W . (g)-(i) The laser power is 1.9 mW with the laser focused at one spot to evaluate the e\u000bect of laser\ndamage.\nnetization is small. We estimate the uncertainty as the di\u000berence between magnetization at 200 Oe and 500 Oe.\n13(a) (b)\n(c) (d)FIG. 9. Magnetization 4 \u0019Mvs. \feld for GE varnish and sample 3. (a)-(b) Quartz sample holder and GE varnish response.\nMeasured moment is normalized to initial V[TCNE] xvolume. (a) Total magnetization (b) Linear background subtracted\nmagnetization. (c)-(d) V[TCNE] xsample on day 3. (c) Total magnetization (d) Linear background subtracted magnetization.\n14" }, { "title": "1706.06379v1.Magnetic_signatures_of_quantum_critical_points_of_the_ferrimagnetic_mixed_spin__1_2__S__Heisenberg_chains_at_finite_temperatures.pdf", "content": "arXiv:1706.06379v1 [cond-mat.stat-mech] 20 Jun 2017Journal of Low Temperature Physics manuscript No.\n(will be inserted by the editor)\nMagnetic signatures of quantum critical points of\nthe ferrimagnetic mixed spin-(1/2, S) Heisenberg\nchains at finite temperatures\nJozef Streˇ cka ·Taras Verkholyak\nReceived: date / Accepted: date\nAbstract Magneticpropertiesoftheferrimagneticmixedspin-(1/2, S)Heisen-\nberg chains are examined using quantum Monte Carlo simulations for t wo dif-\nferent quantum spin numbers S= 1 and 3/2. The calculated magnetization\ncurves at finite temperatures are confronted with zero-temper ature magne-\ntization data obtained within density-matrix renormalization group m ethod,\nwhich imply an existence of two quantum critical points determining a b reak-\ndown of the gapped Lieb-Mattis ferrimagnetic phase and Tomonaga -Luttinger\nspin-liquid phase, respectively. While a square-root behavior of the magneti-\nzation accompanying each quantum critical point is gradually smooth ed upon\nrising temperature, the susceptibility and isothermal entropy cha nge data pro-\nvide a stronger evidence of the quantum critical points at finite tem peratures\nthrough marked local maxima and minima, respectively.\nKeywords ferrimagnetic Heisenberg chains ·quantum critical point ·\nquantum Monte Carlo\nPACS75.10.Pq ·75.10.Kt ·75.30.Kz ·75.40.Cx ·75.60.Ej\nThis work was financially supported by the grant of The Minist ry of Education, Science,\nResearch and Sport of the Slovak Republic under the contract No. VEGA 1/0043/16, as\nwell as, by grants of the Slovak Research and Development Age ncy provided under Contract\nNos. APVV-0097-12 and APVV-14-0073.\nJozef Streˇ cka\nInstitute of Physics, Faculty of Science of P. J. ˇSaf´ arik University,\nPark Angelinum 9, 040 01 Koˇ sice, Slovak Republic\nE-mail: jozef.strecka@upjs.sk\nTaras Verkholyak\nInstitute for Condensed Matter Physics, NASU,\n1 Svientsitskii Street, 790 11 L’viv-11, Ukraine\nE-mail: werch@icmp.lviv.ua2 Jozef Streˇ cka, Taras Verkholyak\n1 Introduction\nQuantum phase transitions traditionally attract a great deal of at tention, be-\ncause they are often accompanied with several remarkable signat ures exper-\nimentally accessible at nonzero temperatures [1]. One-dimensional q uantum\nspin chains provide notable examples of condensed matter systems , which\nbring a deeper understanding into exotic forms of magnetism nearb y quan-\ntum critical points based on rigorous calculations [2]. Apart from con ven-\ntional solid-state compounds affording experimental realizations o f the quan-\ntum spin chains [3], one may simulate a quantum phase transition in spin\nchains through ultracold atoms confined in an optical lattice [4]. Frac tional\nmagnetizationplateaux[5],magnetizationcuspsingularities[6]andsp in-liquid\nground states [5,6,7,8] can be regarded as the most profound m anifestations\nof zero-temperature magnetization curves of the quantum spin c hains.\nThe ferrimagnetic mixed spin-( s,S) Heisenberg chains [9,10,11,12,13,14,\n15,16,17,18] with regularly alternating spins s= 1/2 andS >1/2 display\nthe intermediate magnetization plateau inherent to the gapped Lieb -Mattis\nferrimagnetic ground state, as well as, the gapless Tomonaga-Lu ttinger spin-\nliquid phase. It has been argued [18] on the grounds of Lieb-Mattis t heorem\n[19] and Oshikawa-Yamanaka-Affleck rule [20] that the ferrimagnet ic mixed\nspin-(1/2, S)Heisenberg chains should exhibit just one plateau at the following\nvalue of the total magnetization m/ms= (2S−1)/(2S+1) normalized with\nrespect to its saturation. The gapped Lieb-Mattis ferrimagnetic g round state\nbreaks down at a quantum phase transition towards the gapless To monaga-\nLuttinger spin-liquid phase, which is accompanied with a singular squar e-root\nbehaviorof the magnetization. The same type of the magnetization singularity\nalsoappearsjustbelowthesaturationfield,whichdeterminesaqua ntumphase\ntransition between the Tomonaga-Luttinger spin liquid and the fully p olarized\nphase.Inthepresentworkwewillexaminemagneticsignaturesoft hequantum\ncritical points of the ferrimagnetic mixed spin-(1/2, S) Heisenberg chains at\nfinite temperatures by making use of quantum Monte Carlo simulation s.\n2 Model and method\nLetusconsidertheferrimagneticmixedspin-(1/2, S)Heisenbergchainsdefined\nthrough the Hamiltonian\nˆH=JL/summationdisplay\nj=1ˆSj·(ˆsj+ˆsj+1)−hL/summationdisplay\nj=1(ˆSz\nj+ ˆsz\nj), (1)\nwhereˆsj≡(ˆsx\nj,ˆsy\nj,ˆsz\nj) andˆSj≡(ˆSx\nj,ˆSy\nj,ˆSz\nj) denote the standard spin-1/2\nand spin- Soperators, respectively. The first term in the Hamiltonian (1) take s\ninto account the antiferromagnetic Heisenberg interaction J >0 between the\nnearest-neighborspins and the second term h=gµBHincorporatingthe equalMagnetic signatures of quantum critical points of the Heise nberg chains 3\nLand´ e g-factors gs=gS=gand Bohr magneton µBaccounts for the Zee-\nmann’s energy of individual magnetic moments in an external magnet ic field.\nSince the elementary unit contains two spins an overall chain length is 2L,\nwhereas a translational invariance is ensured by the choice of perio dic bound-\nary conditions ˆsL+1≡ˆs1.\nTo explore magnetic properties of the ferrimagnetic mixed spin-(1/ 2,S)\nHeisenberg chains at nonzero temperatures, we have implemented a directed\nloopalgorithminthestochasticseriesexpansionrepresentationof thequantum\nMonte Carlo (QMC) method [21] from Algorithms and Libraries for Phy sics\nSimulations (ALPS) project [22]. The QMC method allows a straightfor ward\ncalculation of the magnetization data at finite temperatures, which will be\nalso confronted with recent zero-temperature magnetization da ta calculated\nwithin the density-matrix renormalization group (DMRG) method [18] serv-\ning as a useful benchmark at low enough temperatures. The susce ptibility of\nthe ferrimagnetic mixed spin-(1/2, S) Heisenberg chains can also be directly\ncalculated from a directed loop algorithm of QMC method, while the isot her-\nmal entropy change can be obtained from the relevant magnetizat ion data\nusing the Maxwell’s relation\n∆ST=/integraldisplayh\n0/parenleftbigg∂m\n∂T/parenrightbigg\nhdh. (2)\nTo avoid an extrapolation due to finite-size effects, we have perfor med QMC\nsimulations for a sufficiently large system size with up to L= 128 units (256\nspins), whereas adequate numerical accuracy was achieved thro ugh 750 000\nMonte Carlo steps used in addition to 150 000 steps for thermalizatio n.\n3 Results and discussion\nLet us proceed to a discussion of the most interesting results for t he magneti-\nzation and susceptibility data of the ferrimagnetic mixed spin-(1/2, S) Heisen-\nberg chains. Fig. 1(a) shows a three-dimensional (3D) plot of the m agneti-\nzation of the ferrimagnetic mixed spin-(1/2,1) Heisenberg chain aga inst the\nmagnetic field and temperature. As one can see, the one-third plat eau due to\nthe gapped Lieb-Mattis ferrimagnetic ground state diminishes upon increas-\ning of temperature until it becomes completely indiscernible above ce rtain\ntemperature kBT/J≈0.5. To explore the temperature effect in more detail,\nFig. 1(b) compares a zero-temperature magnetization curve obt ained within\nDMRG method [18] with low-temperature magnetization data stemmin g from\nQMC simulations. It is quite evident that the square-root singularity of the\nmagnetization, which emerges at both quantum critical points dete rmining an\nupper edge of the one-third plateau and saturation field, is gradua lly rounded\nupon raising temperature.\nA stronger evidence of two quantum phase transitions of the ferr imagnetic\nmixed spin-(1/2,1)Heisenberg chain thus provides the susceptibility , which is\nplotted in Fig. 2. As a matter of fact, the susceptibility still displays a t low4 Jozef Streˇ cka, Taras Verkholyak\n/s48/s49/s50/s51/s52\n/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s107\n/s66/s32\n/s84\n/s32/s47/s32\n/s74/s109/s32/s47/s32/s109/s115\n/s104/s32/s47/s32/s74/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54/s48/s46/s55/s48/s46/s56/s48/s46/s57/s49/s46/s48/s109 /s32/s47/s32 /s109\n/s115\n/s40/s97/s41\n/s48 /s49 /s50 /s51/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s40/s98/s41/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s48/s32/s40/s68/s77/s82/s71/s41\n/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s49/s32/s32/s40/s81/s77/s67/s41\n/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s50/s32/s32/s40/s81/s77/s67/s41\n/s104\n/s115/s32/s47/s32 /s74/s32 /s61/s32/s51/s46/s48/s48/s104\n/s99/s32/s47/s32 /s74/s32 /s61/s32/s49/s46/s55/s54/s49/s47/s51 /s112/s108/s97/s116/s101/s97/s117/s49/s47/s50 /s49\n/s32/s32/s109/s32/s47/s32/s109\n/s115\n/s104 /s32/s47/s32 /s74\nFig. 1 The total magnetization of the ferrimagnetic mixed spin-(1 /2,1) Heisenberg chain\nnormalized with respect to its saturation value: (a) 3D surf ace plot as a function of temper-\nature and magnetic field constructed from QMC data; (b) Zero- temperature DMRG data\nversus QMC magnetization curves at low enough temperatures .\n/s48\n/s49\n/s50\n/s51\n/s52/s48/s46/s51/s48/s46/s54/s48/s46/s57/s49/s46/s50\n/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56\n/s40/s97/s41\n/s32/s74\n/s107/s66/s32/s84/s32/s47/s32/s74\n/s104\n/s32/s47/s32\n/s74/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s53/s48/s46/s54/s48/s46/s55/s48/s46/s56/s48/s46/s57/s74\n/s48 /s49 /s50 /s51 /s52/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56\n/s40/s98/s41/s107\n/s66/s32/s84 /s32/s47/s32 /s74/s32\n/s32/s48/s46/s48/s32/s40/s68/s77/s82/s71/s41\n/s32/s48/s46/s48/s53/s32/s40/s81/s77/s67/s41\n/s32/s48/s46/s49/s48/s32/s40/s81/s77/s67/s41\n/s32/s48/s46/s50/s48/s32/s40/s81/s77/s67/s41\n/s104\n/s115/s32/s47/s32 /s74/s32 /s61/s32/s51/s46/s48/s48/s104\n/s99/s32/s47/s32 /s74/s32 /s61/s32/s49/s46/s55/s54/s115/s112/s105/s110/s32/s108/s105/s113/s117/s105/s100/s49/s47/s50 /s49\n/s32/s32/s74\n/s104 /s32/s47/s32 /s74\nFig. 2The susceptibility of the ferrimagnetic mixed spin-(1/2,1 ) Heisenberg chain per unit\ncell: (a) 3D surface plot as a function of temperature and mag netic field constructed from\nQMCdata; (b) Zero-temperature DMRGdata versus QMCsimulat ions at lowtemperatures.\nenough temperatures two sharp peaks in a vicinity of the magnetic fi elds driv-\ning the investigated quantum spin system towards the quantum crit ical points\nin addition to a pronounced divergence observable at zero magnetic field. Of\ncourse, these local maxima are gradually suppressed and merge to gether upon\nincreasing of temperature.\nTo illustrate a general validity of the aforedescribed conclusions, F igs. 3\nand 4 show similar plots for the magnetization and susceptibility of ano ther\nferrimagnetic mixed spin-(1/2,3/2) Heisenberg chain. It is quite app arent that\nthe same general trends can be reached as far as the temperatu re effect is con-\ncerned. There are only two quantitative differences. The first diffe rence refers\nto a height of the intermediate Lieb-Mattis plateau, which is in the pre sent\ncase at one-half of the saturation magnetization. The second diffe rence lies\nin absolute values of two field-driven quantum critical points, which d eter-\nmine the end of the intermediate one-half plateau and the beginning o f the\nsaturation magnetization, respectively.Magnetic signatures of quantum critical points of the Heise nberg chains 5\n/s48/s49/s50/s51/s52/s53\n/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s107\n/s66/s32\n/s84\n/s32/s47/s32\n/s74/s109/s32/s47/s32/s109/s115\n/s104/s32/s47/s32/s74/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54/s48/s46/s55/s48/s46/s56/s48/s46/s57/s49/s46/s48/s109 /s32/s47/s32 /s109\n/s115\n/s40/s97/s41\n/s48 /s49 /s50 /s51 /s52/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56/s49/s46/s48\n/s40/s98/s41/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s48/s32/s40/s68/s77/s82/s71/s41\n/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s49/s32/s32/s40/s81/s77/s67/s41\n/s32/s107\n/s66/s84/s32 /s47/s32 /s74 /s32/s61/s32/s48/s46/s50/s32/s32/s40/s81/s77/s67/s41\n/s49/s47/s50 /s112/s108/s97/s116/s101/s97/s117\n/s49/s47/s50 /s51/s47/s50\n/s32/s32/s109/s32/s47/s32/s109\n/s115\n/s104 /s32/s47/s32 /s74/s104\n/s115/s32/s47/s32 /s74/s32 /s61/s32/s52/s46/s48/s48/s104\n/s99/s32/s47/s32 /s74/s32 /s61/s32/s50/s46/s56/s52\nFig. 3The total magnetization of the ferrimagnetic mixed spin-(1 /2,3/2) Heisenberg chain\nnormalized with respect to its saturation value: (a) 3D surf ace plot as a function of temper-\nature and magnetic field constructed from QMC data; (b) Zero- temperature DMRG data\nversus QMC magnetization curves at low enough temperatures .\n/s48\n/s49\n/s50\n/s51\n/s52\n/s53/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54\n/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56\n/s40/s97/s41\n/s32/s74\n/s107/s66/s32/s84/s32/s47/s32/s74\n/s104\n/s32/s47/s32\n/s74/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s52/s48/s46/s53/s74\n/s48/s46/s48 /s48/s46/s53 /s50 /s51 /s52 /s53/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54\n/s40/s98/s41/s107\n/s66/s32/s84 /s32/s47/s32 /s74/s32\n/s32/s48/s46/s48/s32/s40/s68/s77/s82/s71/s41\n/s32/s48/s46/s48/s53/s32/s40/s81/s77/s67/s41\n/s32/s48/s46/s49/s48/s32/s40/s81/s77/s67/s41\n/s32/s48/s46/s50/s48/s32/s40/s81/s77/s67/s41\n/s115/s112/s105/s110/s32/s108/s105/s113/s117/s105/s100/s49/s47/s50 /s51/s47/s50\n/s32/s32/s74\n/s104 /s32/s47/s32 /s74/s104\n/s115/s32/s47/s32 /s74/s32 /s61/s32/s52/s46/s48/s48/s104\n/s99/s32/s47/s32 /s74/s32 /s61/s32/s50/s46/s56/s52\nFig. 4 The susceptibility of the ferrimagnetic mixed spin-(1/2,3 /2) Heisenberg chain per\nunit cell: (a) 3D surface plot as a function of temperature an d magnetic field constructed\nfrom QMC data; (b) Zero-temperature DMRG data versus QMC sim ulations at low enough\ntemperatures.\nLast but not least, let us examine the magnetic-field variations of th e\nisothermal entropy change, which represents one of two basic ma gnetocaloric\npotentials. For this purpose, Fig. 5 illustrates typical dependence s of the\nisothermal entropy change of two considered ferrimagnetic mixed spin-(1/2, S)\nHeisenberg chains as a function of the magnetic field at a few differen t temper-\natures. It is quite clear that the isothermal entropy change −∆STexhibits a\nsteep increase when the magnetic field is lifted from zero, which can b e related\nto an abrupt field-induced increase of the magnetization from zero towardsthe\nferrimagnetic Lieb-Mattis plateau m/ms= (2S−1)/(2S+1). The isothermal\nentropy change −∆STthen reaches a constant value within the field range,\nwhich is nearly equal to a width of the intermediate magnetization plat eau\nat a given temperature. This interval of the magnetic fields determ ines range\nof applicability of the ferrimagnetic mixed spin-(1/2, S) Heisenberg chains for\ncooling purposes. Namely, the isothermal entropy change −∆STconsecutively6 Jozef Streˇ cka, Taras Verkholyak\n/s48 /s49 /s50 /s51 /s52/s45/s48/s46/s48/s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48/s48/s46/s49/s53/s48/s46/s50/s48/s48/s46/s50/s53/s48/s46/s51/s48\n/s48/s46/s50\n/s48/s46/s49\n/s107\n/s66/s32/s84\n/s32/s47/s32 /s74/s32 /s61/s32/s48/s46/s48/s53/s49/s47/s50 /s49\n/s32/s32/s83\n/s84/s32/s47/s32/s107\n/s66\n/s40/s97/s41 /s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s104 /s32/s47/s32 /s74/s48 /s49 /s50 /s51 /s52 /s53/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48/s48/s46/s49/s53/s48/s46/s50/s48/s48/s46/s50/s53/s48/s46/s51/s48\n/s48/s46/s50\n/s48/s46/s49\n/s107\n/s66/s32/s84\n/s32/s47/s32 /s74/s32 /s61/s32/s48/s46/s48/s53/s49/s47/s50 /s51/s47/s50\n/s32/s32/s83\n/s84/s32/s47/s32/s107\n/s66\n/s40/s98/s41 /s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s104 /s32/s47/s32 /s74\nFig. 5 The isothermal change of the entropy versus the magnetic fiel d at three different\ntemperatures for the ferrimagnetic mixed spin-(1/2, S) Heisenberg chain: (a) S= 1; (b)\nS= 3/2.\nacquiresrelativelydeep minima near both field-driven quantum critica lpoints,\nwhich are gradually lifted and smoothed upon increasing of temperat ure. The\nminima in the isothermal entropy change −∆STcan be thus regarded as an-\nother faithful indicators of the quantum critical points of the fer rimagnetic\nmixed spin-(1/2, S) Heisenberg chains.\n4 Conclusion\nIn the present work we have examined magnetic and thermodynamic proper-\ntiesoftheferrimagneticmixedspin-(1/2, S)Heisenbergchainsatfinitetemper-\natures using QMC simulations. In particular, we have focused our at tention to\na detailed examination of the temperature effect upon the magnetiz ation pro-\ncess in a close vicinity of the field-driven quantum critical points. It h as been\nevidenced that a singularsquare-rootbehaviorof the magnetizat ion, which ac-\ncompanies both quantum phase transitions connected with a break downof the\ngapped Lieb-Mattis ferrimagnetic phase and the gapless Tomonaga -Luttinger\nspin-liquid phase, undergoes a gradual rounding upon increasing of tempera-\nture.Themagnetizationcurveatnon-zerotemperaturesisthus almostwithout\nany clear signature of the quantum critical points unlike other ther modynamic\nresponse functions.\nThe susceptibility and isothermal entropy changes contrarily displa y close\nto quantum phase transitions relatively sharp maxima and minima, res pec-\ntively, which are gradually suppressed and smoothed upon increasin g of tem-\nperature. The respective maximum in the susceptibility and minimum in t he\nisothermal entropy change can be accordingly regarded as a faith ful indicator\nof the field-driven quantum critical point of the ferrimagnetic mixed spin-\n(1/2,S) Heisenberg chains. Although the present study was restricted j ust to\nthe mixed-spin Heisenberg chains with the quantum spin numbers S= 1 and\n3/2, the foregoing studies [9,10,11,12,13,14,15,16,17,18] h ave already proven\nlargely universal behavior of the ferrimagnetic mixed spin-(1/2, S) HeisenbergMagnetic signatures of quantum critical points of the Heise nberg chains 7\nchains also for other spin values for which qualitatively the same beha vior\nshould be expected.\nIt has been also demonstrated that the most efficient cooling throu gh the\nadiabatic demagnetization of the ferrimagnetic mixed spin-(1/2, S) Heisenberg\nchains can be achieved in a range of the magnetic fields from zero up t o nearly\na half of the intermediate magnetization plateau, while the isotherma l entropy\nchange implies for larger magnetic fields a less efficient cooling as it rapid ly\ndrops down near both quantum critical points. It is our hope that t heoretical\nresults of the present work are ofobvious relevance for a series o f bimetallic co-\nordinationcompoundsMM’(pba)(H 2O)3·2H2O [23] andMM’(EDTA) ·6H2O\n[24] (M,M’ = Cu, Ni, Co, Mn), which represent experimental realizatio ns of\nthe ferrimagnetic mixed-spin Heisenberg chains. High-field magnetiz ation and\nmagnetocaloric measurements on these or related series of bimeta llic com-\nplexes are however needed for experimental testing of the prese nt theoretical\npredictions.\nReferences\n1. S. Sachdev, Quantum Phase Transitions . Cambridge University Press, Cambridge\n(1999).\n2. D.C. Mattis, The Many-Body Problem: An Encyclopedia of Exactly Solved Mo dels in\nOne Dimension . World Scientific, Singapore (1993).\n3. J.S. Miller, M. Drillon, Magnetism: Molecules to Materials I . Wiley-VCH Verlag, Wein-\nheim (2001).\n4. J. Simon, W.S. Bakr, R. Ma, M.E.Tai, P.M.Preiss,M. Greine r,Nature472,307 (2011).\n5. A. Honecker, J. Schulenburg, J. Richter, J. Phys.: Condens. Matter 16, S749 (2004).\n6. K. Okunishi, Prog. Theor. Phys. Suppl. 145, 119 (2002).\n7. G. Misguich, Quantum Spin Liquids , inExact Methods in Low-Dimensional Statistical\nPhysics and Quantum Computing , eds. J. Jacobsen, S. Ouvry, V. Pasquier, D. Serban,\nL. F. Cugliandolo. Oxford University Press, Oxford (2008).\n8. Y. Zhou, K. Kanoda, T.-K. Ng, preprint arxiv: 1607.03228\n9. T. Kuramoto, J. Phys. Soc. Jpn. ,67, 1762 (1998).\n10. S. Yamamoto, T. Sakai, J. Phys.: Condens. Matter ,11, 5175 (1999).\n11. T. Sakai, S. Yamamoto, Phys. Rev. B ,60, 4053 (1999).\n12. N.B. Ivanov, Phys. Rev. B ,62, 3271 (2000).\n13. A. Honecker, F. Mila, M. Troyer, Eur. Phys. J. B ,15, 227 (2000).\n14. S. Yamamoto, T. Sakai, Phys. Rev. B ,62, 3795 (2000).\n15. T. Sakai, S. Yamamoto, Phys. Rev. B ,65, 214403 (2002).\n16. A.S.F. Ten´ orio, R.R. Montenegro-Filho, M.D. Coutinho -Filho,J. Phys.: Condens. Mat-\nter,23, 506003 (2011).\n17. N.B. Ivanov, S.I. Petrova, J. Schnack, Eur. Phys. J. B ,89, 121 (2016).\n18. J. Streˇ cka, preprint arxiv: 1607.03617.\n19. E. Lieb, D. Mattis, J. Math. Phys. ,3, 749 (1962).\n20. M. Oshikawa, M. Yamanaka, I. Affleck, Phys. Rev. Lett. ,78, 1984 (1997).\n21. A.W. Sandvik, Phys. Rev. B 59, 14157 (1999).\n22. B. Bauer, L.D. Carr, H.G. Evertz, A. Feiguin, J. Freire, S . Fuchs, L. Gamper, J. Gukel-\nberger, E. Gull, S. Guertler, A. Hehn, R. Igarashi, S.V. Isak ov, D. Koop, P.N. Ma, P.\nMates, H. Matsuo, O. Parcollet, G. Pawlowski, J.D. Picon, L. Pollet, E. Santos, V.W.\nScarola, U. Schollw¨ ock, C. Silva, B. Surer, S. Todo, S. Treb st, M. Troyer, M.L. Wall, P.\nWerner, S. Wessel, J. Stat. Mech.: Theor. Exp. ,2011, P05001 (2011).\n23. O. Kahn, Struct. Bonding (Berlin) ,68, 89 (1987).\n24. M. Drillon, E. Coronado, D. Beltran, R. Georges, J. Appl. Phys. ,57, 3353 (1985)." }, { "title": "2105.10795v1.Ferromagnetic_resonance_modes_in_the_exchange_dominated_limit_in_cylinders_of_finite_length.pdf", "content": " 1 Ferromagnetic resonance modes in the exchange dominated limit \nin cylinders of finite length \n \nJinho Lim1, Anupam Garg1 and J. B. Ketterson1,2 \n \n1. Department of Physics and Astronomy, Northwestern University, Evanston, IL, \n60208. \n2. Department of Electrical and C omputer Engineering Northwestern University, \nEvanston, IL, 60208. \nAbstract \nWe analyze the magnetic mode structure of axially -magnetized, finite -length, nanos copic \ncylinders in a regime where the exchange interaction dominates, along with simulations of the \nmode frequencies of the ferrimagnet yttrium iron garnet. For the bulk modes we find that the \nfrequencies can be represented by an expression given by Herring and Kittel by using \nwavevector components obtained by fitting the mode patterns emerging from the se simulations. \nIn addition to the axial, radial, and azimuthal modes that are present in an infinite cylinder, we \nfind localized “cap modes” that are “trapped” at the top and bottom cylinder faces by the \ninhomogeneous dipole field emerging from the ends. Semi -quantitative explanations are given \nfor some of the modes in terms of a one -dimensional Schrodinger equation which is valid in the \nexchange dominant case. The assignment of the azimuthal mode number is carefully discussed \nand the frequency splitting o f a few pairs of nearly degenerate modes is determined through the \nbeat pattern emerging from them. \n \n1. Introduction \n Measurements of the resonant microwave response of the spins of unpaired electrons in \nferromagnetic materials have a long history, beginni ng with its initial observation by Griffiths1. \nThe vast majority of the subsequent experimental studies have involved “macroscopic” samples \nwhere the spins excited by an external microwave field precess in the combined field arising \nfrom an externally appl ied constant field together with the field produced by the sample’s own \nmagnetization (the so -called demagnetization field) under conditions where both are nearly \nuniform. Under such conditions the strength of the exchange interaction, though itself \nrespon sible for the materials magnetization, does not affect the resonance frequency. \n In the decade after Griffiths’ work, it was found that attempts to drive the resonance to \nlarger amplitudes did not work, and that it saturated at values of the microwave pow er far below \nthat expected from the observed resonance linewidth. The explanation is that small \ninhomogeneities in the magnetization of the sample are amplified by the inhomogeneities \ngenerated in turn in the dipolar or demagnetizing field, and leads to wh at are now called the Suhl \ninstabilities2,3. In addition to studies of the uniformly precessing mode, spatially -nonuniform 2 modes have also been intensively studied, both experimentally and theoretically, as we shall \ndescribe below. \n In this paper, we repor t on an exhaustive numerical study, using the OOMMF \nmicromagnetic simulation code4, of the resonance modes of Yttrium Iron Garnet (YIG) \ncylinders, primarily of diameter d = 75nm and height h = 300 nm, although some aspects have \nbeen studied for other value s of h (7.5 –1200 nm). Our methodology shares many features with \nthe work of McMichael and Stiles5 on two -dimensional elliptical disks and three dimensional \nthin cylindrical discs. Our three -dimensional geometry displays a much richer mode structure, \nhoweve r, requiring a more detailed theoretical framework. In addition, we have also developed \ntechniques to resolve modes that are nearly degenerate in frequency. \n The motivation for our work is as follows. Firstly, in recent years many authors have \nexamined whe ther the magnetization of small particles (in the size range of order 102 nm) can be \nreversed by applying microwave fields (so -called microwave assisted switching) with a view to \napplications in magnetic storage6,7,8,9. The reasoning is that at small sizes , the exchange \ninteraction favors parallel alignment of the spins and suppresses the dipolar instabilities. In \nprevious work, we have used OOMMF to simulate YIG cylinders with (d, h) = (25, 50)nm, \n(25,100)nm and (75,150)nm, where we have attempted to rever se the magnetic moment of the \ncylinder by applying pi -pulses10,11, as in NMR12. We found that by chirping the frequency of the \npulse, we could achieve reversal in the 25 nm diameter samples, but not in the larger ones. If pi -\npulses or other reversal protocols are to be successfully implemented in more realistic larger \nsamples, then an understanding of the normal modes in restricted geometries is a necessary prior \nstep. The problem is also of intrinsic interest, and we have found novel features not seen in \nellipsoidal geometries (including degnerate forms thereof) that have been the subject of almost \nall other work to date. \n Secondly, with recent advances in sub -micron patterning techniques, the study of arrays \nof objects (so as to have large signals) for which the largest dimension is few hundred \nnanometers or less is attracting increasing attention. With advanced techniques it is even possible \nto probe the magnetic properties of individual sub -micron particles13,14,15. Measurements on such \nsamples can even be perf ormed in the absence of an external field, i.e., solely in the presence of \nthe internal demagnetization field (for shapes where such a field exists), provided the sample is \nsmall enough to be in a single domain state16. Modes with an odd number of maxima an d minima 3 can be excited directly with a uniform microwave field; coupling to modes with higher wave \nnumbers will be more challenging17. \n Here we will primarily be concerned with size-quantization effects arising from finite \nsample dimensions. In particular we will examine the mode spectrum in samples having a \ncylindrical shape with radius a (and corresponding diameter d = 2a) and height h, both \nanalytically and numerically. Due to the ease of preparation of some materials as wires, such \nsamples are widely s tudied experimentally, e.g., in permalloy Py18 and in Ni19,20. Cylinders of \nfinite length with h/d ratios of order unity and larger can be readily patterned using optical and e -\nbeam lithography by creating hole arrays in a resist followed by deposition and l iftoff21. \n1.1. Theoretical background \n Free spins in a magnetic field H precess at the Larmor frequency, \nH = , where \ng | e | /2mc=\n with g, e, and m being the electron g -factor, charge and mass. As noted, in \nmaterials having an internal magnetization additional fields are present which can alter the \nprecession frequency. To describe this and related effects, Landau and Lifshitz22 (LL) introduced \nthe following equation of motion \n \n= − − \n0d()dt MMM H M M H ; (1.1) \nhere \nH is the total field at a given position within the sample arising from the external field as \nwell as that produced by the magnetization itself and an effective field arising from quantum \nmechanical exchange; it can also includ e crystalline anisotropy, but this is suppressed in what \nfollows. The second term on the right -hand side of Eq. (1) is incorporated to phenomenologically \naccount for damping, which will largely be neglected in what follows. In addition to satisfying \nEq. (1 ) \nand MH must satisfy appropriate boundary conditions at the surface of the body. \n For ellipsoidal samples (including degenerate forms thereof), and in the presence of a \nhomogeneous external field \nH , the magn etization \nM is nominally homogeneous as is the \nresulting demagnetization field; one can then observe sharp absorption lines in ferrommagnetic \nresonance (FMR) experiments (in the absence of strong damping), all spins then seeing th e same \nlocal field. The resonance frequency of this uniformly precessing mode in a spheroidal sample \n(where two of the principal axes of the ellipsoid are identical) with the external field \n0H along \nthe axis of rotation is given by what is commonly called the Kittel formula,23 4 \n( ) 00H 4 (N N )M⊥ = + −\n , (1.2) \nwhere \nN and N⊥\n are coefficients accounting for the effect of demagnetization perpendicular \nand parallel to the rotation axes (with \n2N N 1⊥+=\n ), and \n0M is the internal magnetization, \ntaken as a constant; note \n may differ from the free -space value due to atomic and solid -state \neffects. \n In addition to the uniformly precessing mode ther e exist non -uniform modes24 which we \ncan characterize by some effective wavelength, \n . At shorter (nanometer) scale wavelengths, the \nexchange interaction dominates, and the associated modes are termed exchange modes, first \nintroduc ed by Bloch25. The importance of modes with longer wavelengths (in suitably large \nsamples), was emphasized by Clogston, Suhl, Walker and Anderson26,27. They arise from a \nsolution of Eq. (1) together with \n0 =B and the Maxwell boundary co nditions; they are \ncommonly referred to as magnetistatic modes . Modes in the region where both exchange and \nmagnetostic effects compete are called dipole/exchange modes. \n For the case of a sphere some of the low -lying magnetostaic modes were first examined \nby Mercerau and Feynman28. They were later studied in much greater detail for spheroidal \nsamples by Walker29. \n The limiting case of a finite thickness, infinite -area, slab (\nz N 0, N 4⊥= = ) with both \nthe wave vector \nk (\n| k | 2 /= ) and the external magnetic field \n0H lying in plane was treated \nby Damon and Eshbach30; the case with \n0H perpendicular to the film was examined by Damon \nand Van De Vaart31. For an in-plane field Damon and Eshbach identified two classes of \nmagnetostatic modes: a surface wave and a family of bulk waves. The first of these, now \ndesignated as the DE mode, decays exponentially within the interior. The DE mode is most \ncommonly studied for wave vectors \n0ˆ () k H z\n where \nˆz is the plane normal; it has a positive \ngroup velocity but also has the unusual property that it propagates only on one side of the slab \nfor a given field direction, switching to th e opposite side on reversing the field. The second class \nis a family of bulk modes which are quantized in the film thickness direction; they are typically \nstudied with \n0 kH\n . They propagate in both directions and the lowest lying mode has the \nunusual property that its group velocity is negative for small k and therefore it is called a \nbackward volume mode . The frequencies of all the volume modes asymptotically approach the 5 free spin precession frequency, \n0H = , as \nk→ (in the absence of exchange); they approach \nthe in -plane Kittel frequency, \n0 0 0(H (H 4 M ) + , as \nk0→ . \n For \n0H perpendicular to the film we again have modes quantized a long the film \nthickness that now propagate isotropically in plane. The lowest has a positive group velocity for \nsmall in -plane wave vector k; it is therefore referred to as a forward volume mode . All these \nmodes approach the perpendicular Kittel frequency \n0(H 4 M) − as \nk0→ and \n0 0 0(H (H 4 M ) − \n as \nk→ . Exchange effects have been considered by De Wames and \nWolfram32 and more completely by Arias33. \n For the case of an infinitely long cyl inder (\nz N 2 ;N 0⊥= = ) with \n0H parallel to the \nrotational axis, which is relevant to the work presented here, the mode structure was first studied \nby Joseph and Schlomann34. Here we encounter families of purely azimutha l modes and \nanalogous to the backward volume modes we have radially quantized modes propagating up and \ndown the cylinder axis, which also approach \n0H = at large k (in the absence of exchange). \nRecently this problem was reexamined by Arias and Mills35 who also considered the effects of \nexchange via perturbation theory. \n At shorter wavelengths the effects of exchange contribute. In this regime the frequency of \na mode with wave vector \nk 2 /= for a spheroidal sample wi th the external field \n0H aligned \nalong the rotational axis can be described by the Herring -Kittel formula,36 which we discuss in \nAppendix A \n \n2 2 20 0 ex 0 0 ex 0(H 4 N M D k )(H 4 N M D k 4 M sin ) = − + − + + \n ; (1.3) \nhere \nexD is a parameter measuri ng the strength of the exchange (see below), \n2 2 2z k k k⊥=+ , \nzk and k⊥\n are the components of the wavevector parallel and perpendicular to the spheroid axis, \nand \n1z tan (k / k )−⊥ = is the angle between the spin wave p ropagation direction and the \nspheroid axes. Note that at k = 0 the factor involving \nN⊥ that appears in Eq. (1.2) is absent from \nEq. (1.3), since it is assumed that the transverse demagnetization field is “screened out” at short \nwavelengths. Indeed \n is ill defined at precisely k = 0 since \n is ambiguous. This shortcoming \nalso signals the importance of the magnetostatic modes at intermediate k values; i.e., as the \nsample size is red uced there is a crossover between dipole dominated and exchange dominated 6 modes. Modes with k values straddling these regimes are the dipole -exchange modes mentioned \nabove. At short wavelengths, which will be the case in sufficiently small samples, Eq. (1. 3) \nshould provide a representation of the mode structure in rotationally symmetric samples, \nprovided quantized values of \nz k and k⊥ satisfying the boundary conditions are available; we \nwill utilize Eq. (1.3) to represent some of our finite -size sample simulations in what follows. \n More generally and in the absence of exchange effects, magnetostatic effects would \ndominate the mode frequencies, which for a spheroid would lie in the range \n \n0 0 0 0 0(H 4 N M 2 M ) (H 4 N M ) − + − \n ; (1.4) \nnote the numbe r of modes in this interval is bounded only by the number of spins; i.e., the mode \ndensity is very high, making the resolution of the individual modes extremely difficult at shorter \nwavelengths (where they pile up). When exchange is present the mode freque ncies are spread \nover a much wider interval. \n In an inhomogeneous external field, or for samples with an arbitrary shape, one might \ninitially expect to observe a line -broadening (as happens in most nuclear magnetic resonance \nexperiments). However, in the p resence of exchange this is not the case and well-defined modes \nemerge as will be discussed further below. Exchange modes are not restricted to magnetic \nmaterials; they have been observed in experiments on Fermi liquids at low temperatures, \nexamples being certain metals37 and the normal state of liquid 3He 38; the theory for the latter was \nfirst given by Silin39 for homogeneous fields and later extended by Leggett40 to describe the \ninhomogeneous case. \n1.2. Plan of the paper \n We develop the theoretical framewor k for our problem in Sec. 2, beginning with a \ndiscussion of cylindrical symmetry and the resulting angular momentum quantum number (or \nazimuthal mode number) in Sec. 2.1. In the magnetostatic limit it is convenient to take this as the \ntotal angular momentu m, m, as done by Walker, and by Joseph and Schlomann34. In the \nexchange dominated limit, it becomes more important to understand the separation of the angular \nmomentum into its orbital and spin parts. The major component of a mode has spin \nsm1= and \norbital angular momentum \nmpl= , and there is a small admixture of \nsm1=− and \nm p 2l=+ . \nAccordingly, we find it better to label the modes by the orbital angular momentum p of the major 7 component. This is especially so when examining the computer -generated mode patterns since \nthe orbital behavior of any component of \nM is immediately apparent. \n That the two components of a mode are so unequal goes hand in ha nd with the fact that \nmodes with m l = p and −p are nearly degenerate, as we discuss in Sec. 2.2. A clear understanding \nof this issue is important, as this near degeneracy can lead to some confusion when looking at the \nmode patterns. In one case, we have re solved this degeneracy (see Sec. 3. 4) by exciting and \nexamining the beat pattern between the ±p modes. \n In Sec. 2.3, we show that exchange dominated modes in long cylinders are approximately \ndescribed via a Schrodinger -like equation for a particle in a cylindrical box with a modified \nbounda ry condition, such that axial and radial dependence of the mode function factorizes, and \nthe resulting quantization gives rise to axial and radial mode numbers. In an infinite cylinder this \nseparation is exact, which is exploited to good effect in the anal yses of Joseph and Schlomann34 \nand of Arias and Mills35. In a finite cylinder, the separation is approximate since the \ndemagnetizing field is non -uniform, and flares away from the axis near the perimeter of caps at z \n= 0 and z = h. We give a semi -quantitat ive argument in Sec. 2.4 that the Schrodinger equation \npossesses bound state solutions near these caps, corresponding to “cap modes” which we see \nvery clearly in our simulations. For any p, there are two such modes (one for each cap), whose \nfrequencies lie below those of the bulk modes with the same p. This means that the uniform \nFMR or Kittel mode, which is the lowest bulk mode with p = 0, is not the lowest frequency mode \nof the body. For this case, we present numerical results for the solution of the Schr odinger \nequation in Sec. 2.5, and find good agreement with the simulations. \n The cap modes are a novel and unexpected feature of our study, as they do not exist in an \ninfinite cylinder or a finite sized ellipsoid of revolution. Similar “end modes” were fou nd by \nMcMichael and Stiles5, who did not however investigate their origin. We expect that such \nlocalized modes will exist near the surfaces of other sample shapes as well whenever the \ndemagnetizing field departs significantly from uniformity. \n Our simulati onal approach is described in Sec. 3. It is based on the OOMMF code \ndeveloped at the National Institute of Standards and Technology. After finding the static \nequilibrium magnetization, \neq()Mr (Sec. 3.1), we can excite the system by a pplying pulses that \nare localized in either space, time or both41. The spatial center and the width and frequency \nbandwidth are varied depending on which mode(s) we wish to excite. The resulting time 8 development of \n( ,t)Mr is Fourier tra nsformed, and the point -wise power spectrum is added over \nall the cells. The resulting sum displays peaks at many mode frequencies, and by honing in on \nindividual peaks, we can identify the magnetization patterns for each mode as explained in Sec. \n3.3. \n Once a particular mode pattern is obtained in a simulation it can be used as is or altered in \nsome way, say by combining it with some other mode, to study the subsequent development in \ntime. This is a potentially promising way to study mode -mode coupling or large amplitude \nresponses which we hope to pursue in the future. As an application of this idea, and as noted \nabove, ±p modes are sometimes nearly degenerate, as are the even and odd super -positions of the \ncap modes. In Sec. 3.4 we show that by starting th e simulation in a suitable real -space pattern we \ncan find a beat pattern in the time development of the magnetization from which we can obtain \nthe frequency splitting of the modes. We have performed this exercise for only a few cases as it \nis computational ly intensive, and the physical principles are the same for the other cases. \n In Sec. 4 we tabulate the frequencies of all the modes we have found (approximately 90) \nand discuss the assignment of mode numbers further. The assignment of the longitudinal \nquantum number \nzn on the basis of the one -dimensional Schrodinger equation is particularly \ntricky as the existence of the cap modes forces nodes in the bulk mode functions near the caps, \nand prevents accurate fitting of the lowest few bulk modes to a sinusoidal form \nz sin(k z) with \nzk \nstrictly equal to π/h times an integer. Nevertheless, an unambiguous labeling of the modes is \npossible. \n In Sec. 5 we show that the mode frequencies that we obtain agree surprisingly well with \nthe Herring -Kittel expression (1.3) provided we identify \nk⊥ and \nzk in this formula correctly. \nWe give reasons why this agreement might be so good, explain how the wavevector components \nare found, and how this allows us to organize the normal mode spectrum into familie s of modes \nlabeled by p. \n Spatial patterns for a variety of modes are given in Sec’s. 6 and 7 (in Figs. 6.1 –6.8 and \nFig. 7.2). These patterns are the centerpiece of our paper, and show beautiful regularity and \nsymmetry. In Sec. 6 we consider only the d = 7 5 nm, h = 300 nm sample, while in Sec. 7 we \nconsider the lowest three p = 0 modes as a function of h. We find that at small h/d (disc -like \nsample), the symmetric cap mode (which has the lowest frequency of all three) is in fact the \nmode that one would rega rd as the uniform FMR or Kittel mode, and its frequency is well fit by 9 the Kittel formula with an appropriate choice of demagnetization coefficients. For large h/d \nhowever, it is the lowest bulk mode (whose frequency lies above the two cap modes) that shou ld \nbe identified with the Kittel mode. For intermediate values of h/d ≃ 6–8, the Kittel formula does \nnot actually describe any of the modes. To our knowledge this point has not been appreciated \nbefore. Once again it illustrates the richness of the normal m ode spectrum in non-ellipsoidal \nsamples. \n Finally, Sec. 8 summarizes our conclusions. Here we take the opportunity to emphasize \nthe importance that simulations of the small amplitude mode structure in nano -structures have \nfor present and possible future ap plications, some of which are currently speculative in character. \n2. Modes in the exchange dominated limit \n In the presence of an isotropic exchange interaction, and neglecting the effects of damping, \nEq. (1.1) takes the form42 \n \n\n= − − \n= − − ex 2\n2\n0\nex 2\n0d 2A\ndt M\nD ,MMM H M\nM H M (2.1) \nwhere \nexA is a parameter fixing the strength of the exchange interaction and \nex ex 0D 2A / M \nHere \nH is the applied magnetic field, \n0ˆHz , plus the dipolar o r demagnetizing field generated by \nM\n. In a cylinder of finite height, the dipolar field is not uniform, especially near the caps, and \nso the static equilibrium field, \neq()Mr is not everywhere parallel to \nˆz . A linearized normal \nmode analysis should therefore consider deviations \neq ( ,t)⊥M r M , which do not lie in the x -y \nplane. If exchange is strong, however, the non -uniformity in \neqM is very small (this is true for \nall the simulations we have performed), and we may then take \nz0 =M . This assumption makes \nit much easier to discuss the physics, and relaxing it only obscures the key ideas without adding \nsubstance. We stress that it is not essential to our argument, especially with respect to the \nsymmetries and the azimuthal quantum number. With this assumption, we may write \n \n( )2 1/20 ˆ ( ,t) M (1 m ) ( ,t) = − + M r z m r , (2.2) \nwhere \nm has only x and y components and is dimensionless since we have scaled out \n0M . \n For small deviations, \n| | 1m , the linearized LL equation can be cast as 10 \n( ) = − − 2\nz 0 d exd ( ,t)ˆH ( ,z) ( ,t) M ( ,t) D ( ,t)dtmrz r m r h r m r . (2.3) \nHere \nzH ( ,z)r consists of the app lied field, \n0ˆHz , together with the position dependent \nlongitudinal demagnetization field arising from the static magnetization, and \ndh is the (small) \ndemagnetization field induced by \nm . (We use cylindrical coordinates \n=(r, ,z)r here and \nbelow.) \n2.1 Assignment of the angular momentum quantum number \n Equation (2.3) defines an eigenvalue problem with cylindrical symmetry, so there must \nexist solutions with definite a zimuthal mode \nnumber. In the zero -exchange or magnetostatic limit, \nthe analysis is best done in terms of a scalar \nmagnetic potential ψ, which varies as \nime in the \neigenmodes; the integer m (which we must be \ncareful to distinguish fr om the scalar value of \nm ) is \nthen naturally interpreted as the angular momentum \nquantum number. In the strong exchange limit, the \nproblem is better formulated in terms of \nm directly, \nwhich as a vector field transforms differently under rotations than a scalar field43 (such as \n()r \nin the Schrodinger equation). \n Let us examine the effect of a rotation on the \nvector \nm at a point \n(x,y,z) by an angle \n about the z axis to a vector \nm at the point \n(x ,y ,z) . \nWe need only carry this analysis out to leading order in \n . The components of the rotated vect or \nm\n are then (see Fig. 2.1.) \n \n = − x x ym (x ,y ,z) m (x,y,z) m (x,y,z) , (2.4a) \n \n = +y x ym (x ,y ,z) m (x,y,z) m (x,y,z) . (2.4b) \nThe coordinates \n(x ,y ,z) themselves are related to \n(x,y,z) as \n \n= + = − x xcos ysin x y , (2.5a) \n \n= + = + y ycos xsin y x . (2.5b) \nIf we expand the left side of (2.4a,b) to first order in \n , note that to zeroth order \nFig. 2. 1. Transf ormation of a vector field. 11 \n = (x ,y ,z) (x,y,z)mm, and recall the definition of the (di mensionless) orbital angular \nmomentum operator \nzl in quantum mechanics as \n \n = − − = − z i x y iyxl , (2.6) \nwe can write \nm in terms of \nm as \n \n = − − −x z x ym (x,y,z) (1 i )m (x,y,z) i [ im (x,y,z)] l , (2.7a) \n \n = − − y z y xm (x,y,z) (1 i )m (x,y,z) i [im (x,y,z)] l . (2.7b) \nEquation (2.7) can be rewritten in the form \n \n − − = − − x x x z\ny y y zm m m 1 i 0 0iim m m 0 1 i i0l\nl . (2.8) \n For a scalar field, \n()r , we would simply have \n = − z ( ) (1 i ) ( )rr l , but the presenc e of \nthe last term in (2.8) mixes the two components of the vector field \nm . This can be interpreted as \narising from an “internal” or “spin” angular momentum of \n=sm1 that is added to or subtracted \nfrom the o rbital angular momentum \nml associated with our vector field \nm (a tensor of rank 1). \nA similar separation exists in the description of light fields44. \n Consider the case of a vector field of the form \n \n=ip\nxm ( ) a(r,z)er , \n=ip\nym ( ) b(r,z)er (a, b arbitrary). (2.9) \nThis field has orbital angular momentum \nmpl , but has no definite spin angular momentum. \nFor it to have a definite spin a and b must be proportion al according to \n \n = x\nsym 1m1m i or \n = − − x\nsym 1m1m i . (2.10a, b) \nWriting \n=xy m m im it then follows that \n \n\n+− = = = +ip\ns totm 1yields m 0, m e and m p 1 (2.11a) \nand \n \n\n+− = − = = −ip\ns totm 1yields m e , m 0 and m p 1 (2.11b) \nwhere we wrote \n=+tot sm m ml . It follows that an eigen mode with total angular momentum \np1+\n must be of the form 12 \n− + −\n−+ =+ − x i(p t) i((p 2) t)\nym ( , t) 11F (r,z) e F (r,z) em ( , t) iir\nr (2.12) \nwhere we have adopted an \n−ite time dependenc e following standard practice. The physical \nsolution is obtained by taking the real part of this complex -valued solution. We shall see below \nthat for positive frequency solutions, \n−+FF in the strong exchange limit, and it is often \nconvenient to neglect \n+F entirely. It is then more useful to label the modes by p, the orbital \nangular momentum of the dominant component, \n−F . This is especially so when looking at mode \npatterns generated by OOMMF, since we can read off p by seeing how many times \nm turns as \nwe go around a circle in the x -y plane. For example, in a \np = 0 mode (see Figs. 6.1a and 6.1b) \nm\n appears uniform, while in p = −1 (Fig. 6.3) and p = 1 (Fig. 6.4) modes \nm winds by 2π and \n−2π, respectively as we go anti -clockwise around a circle. \n In the magnetostatic limit by contrast, \n+− F / F O(1) , and the m labeling is better. Thus, \nfor the sphere, while we would describe the uniform or Kittel mode as having p = 0, Walker29 \nassigns m = 1 to it (see his Fig. 3 where the mode is labeled (110)). \n2.2. Near degeneracy of p and − p modes \n When exchange dominates over dipole -dipole i nteractions, we may as a first approximation \nneglect \ndh in Eq. (2.3). In component form the equation then reads \n \n( )x y 2ex\ny xm ( ,t) m ( ,t) dH(r,z) Dm ( ,t) m ( ,t) dt− = − r r\nr r (2.13) \nor \n \n( )2exdm ( ,t) H(r,z) D m ( ,t)dt = − rr (2.14) \nwith \n \nxy m ( ,t) m ( ,t) im ( ,t) =r r r . (2.15) \nWe seek solutions of the form \n \nitm ( ,t) m ( )e− =rr (2.16) \nand demand that \n0 , with the understanding that the physical solution will be given \nby the real part. These solutions then obey 13 \n( )2z exH (r,z) D m ( ) m ( ) − = \n rr . (2.17) \nThis equation is like a one -particle Schrodinger equation, and since γ > 0 in our convention, the \noperator on the left is a positive operator which cannot have negative eigenvalues. Since we also \ndemand that ω > 0, we must choose \nm0+= . Finally, since our finite cylinder retains full \nazimuthal symmetry, the solution for \nm− takes the form \n \nipm (r, ,z) F (r,z)e−−= , (2.18) \nwhere \nF (r,z)− can be chos en to be real, and ω has the same (positive) value for either sign of p. \nIn terms of the general form (2.12), this solution corresponds to putting \nF0+= , and a physical \nsolution \n \nx i(p t)\nym ( ,t) 1 cos(p t)F (r,z) e c.c. 2F (r,z)m ( ,t) i sin(p t)−−− − = + = − r\nr . (2.19) \n If we now inclu de the dipolar field \ndh as a perturbation, we can expect that \nF+ will \nbecome nonzero, with \n20 ex F / F 4 M / D k+− \n , where \n2 / k is the typical length scale on \nwhich the solution va ries. \n The source of the degeneracy with respect to ±p is that Eq. (2.13) is invariant under \nreflection in the yz plane provided we do not also reflect the vector \nm . Hence the operation \n \nxxm (x,y,z) m ( x,y,z) →− (2.20 a) \nand \n \nyym (x,y,z) m ( x,y,z) →− (2.20b) \nalso produces a solution. This operation is equivalent to φ → − φ, or alternatively to p → − p. \nInclusion of the dipole -dipole interaction destroys this invariance: the field \ndh produced by the \noperation is not the same field as before. \n Strictly speaking therefore, modes differing only in the sign of p are not degenerate, \nalthough the non -degeneracy may be small. Indeed, as explained below, we have spent \nsignificant effort to n umerically resolve the splitting and have not always succeeded. The \nphysical origin of this non -degeneracy is just that the applied external field breaks time reversal \nand parity symmetries. In the magnetostatic limit, this point emerges directly from the solution in \nterms of the scalar potential. Joseph and Schlomann45 find that \n|m| |m|− for the volume 14 modes (where we have here used m instead of p to label the modes), and that only m > 0 \nsolutions exist for the surface modes. This is an e xtreme form of the nondegeneracy, and is the \ncylindrical analog of Damon and Eshbach’s discovery of one -sided surface modes in the slab \ngeometry. Joseph and Schlomann34 also find that the ±p splitting becomes smaller with \nincreasing \nzk or increasing radial mode number (see their Fig. 5). The same behavior is found \nfor the general spheroid by Walker,46. Arias and Mills35 on the other hand, appear to us to be \nfinding that modes with the opposite sign of the angular momentum are degener ate; we are \nunable to pinpoint why. \n The near degeneracy of ±p modes also underlies whether one sees azimuthal standing or \nrunning wave patterns in the OOMMF simulations. We discuss this issue in Sec. 3.3 below. \n2.3. The long cylinder in the exchange domi nated approximation \n For an infinite cylinder the variables in the Schrodinger equation separate and we can \nwrite \n \n( ) ( ) 0 p z z F (r,z) m J (k r) Acos k (z h / 2) Bsin k (z h / 2)−⊥ = − + − (2.21) \nyielding \n \n( ) ( )x\n0 p z z\nym ( ,t) cos(p t)m J (k r) Acos k (z h / 2) Bsin k (z h / 2)m ( ,t) sin(p t)⊥ − = − + − − − r\nr\n . (2.22) \nHere \npJ is the Bessel function of order p. For a long but finite cylinder Eq. (2.21) should be a \ngood approximation except for the cap modes. \n If we take the modes \np+ and \np− as degenerate we can superimpose them and form \nstand ing waves in \n , an operation we carry out in the next section. Inserting any of these forms \ninto (2.8) yields the frequencies \n \n( )22\n0 ex zH D (k k )⊥ = + + . (2.23) \nIf we adopt the boundary condition (discussed below) \n \nd0d=nmr (2.24) \nwhere \nn is a vector normal to the surface, the values of \nk⊥ will be fixed by the condition 15 \npdJ (k a)0dk⊥\n⊥= (2.25) \nwhere a is the cylinder radiu s. We write the solutions of Eq. (2.25) as \nrp,nk where \nrn denotes \nthe number of additional zeros of \npJ (other than those for \np0J at r = 0) within the cylinder o f \nradius a. We will find that (2.25) agrees quite well with the simulations. \n For the finite cylinder we present the argument in two stages. In the first stage we assume \nthat the inhomogenity in the static demagnetization field can be ignored and take \n00 H(r,z) H 4 N M = − \n with \nN\n being the longitudinal demagnetization coefficient. The \nsolution (2.21) continues to hold but the mode frequencies are given by \n \n( )22z 0 0 ex z(k ,k ) H 4 N M D (k k ) ⊥ ⊥ = − + +\n . (2.26) \nThe allowed values of \nk⊥ are given by \nrp,nk as discussed above, but the quantization of \nzk is \nless simple. If the end caps are taken to be at \nz 0 and z h= = , then to have a definite parity under \nreflec tion in the mid plane at \nz h / 2= , the mode function must depend on z as either \n( ) z cos k (z h / 2) −\n (even parity) or \n( ) z sin k (z h / 2) − (odd parity), but the association between \nzk\n and the parity depends on the boundary condition applied at the caps. \n If the boundary condition is taken as \n()0=nm , then the allowed \nzk values are \n \nz z zk , 0, 1, 2h= = . (2.27) \nEven parity is associated with e ven \nz and odd parity with odd \nz . \n If instead the boundary condition is taken as \n0=m , the allowed values of \nzk are \n \nz z zk , 1, 2h= = (2.28 ) \nNow even parity is associated with odd \nz and odd parity with even \nz . \n The boundary condition obeyed by OOMMF mode functions is closer to \n()0=nm \nthan to \n0=m . In addition, they do have definite parity. Except for the two lowest frequency \nmodes, which we call “cap modes” and which require a separate discussion, they are well fit by \nthe \n( ) z cos k (z h / 2) − and \n( ) z sin k (z h / 2) − forms. However, i t is advantageous to allow for a \nshift and write 16 \nzzkh = (2.29) \nwhere \n \nz z z = + \nz( 2) (2.30) \nWe can refer to \nz as an “end defect” analogous to the conc ept of a quantum defect in atomic \nspectroscopy47. With this correction Eq. (2.26) continues to be a good approximation to the mode \nfrequencies. \n We comment further on the boundary condition (2.24) that the normal derivative vanishes \nat the surface. The iso tropic continuum exchange field \n2exDM− arises from a microscopic \nijSS\n Heisenberg interaction, which has the property that for any pair of spins, the torque on \niS \ndue to \njS cancels that on \njS due to \niS . Thus, the total exchange torque on the body vanishes and \nEq. (2.24) is the continuum expression of this fact. This argument dates back to Ament and \nRado48 and has been used by many authors since. Aharoni49 offers a different derivation. Thus, \nit would appear to be very general, and valid for any \nexD , however small. For the magnetostatic \nlimit, \nexD0= , there is however no such condition on \nM . Turning on \nexD perturbatively would \nthen appear to lead to a contradiction. This is not so for the following reason. \n In the boundary value problem for the spatial form of the eigenmodes, \nexD multiplies the \nhighest derivative, and is thus a singular perturbation from the mathematical point of view. Such \nperturbations are known to lead to thin boundary layers where the solution changes character \nrapidly50. Thus, while t he normal derivative at the surface may formally be zero, there could be \nlarge curvature in the boundary layer, and the derivative of \nm as we approach this layer need not \nbe small. This is especially relevant for our OOMMF simulat ions, where the discretization into \ncells may: (a) be too coarse to reveal any boundary layer behavior, and (b) fundamentally \npreclude measurements of this derivative by fitting to the mode functions. In this case adoption \nof an end defect \nz is an effective practical procedure. \n2.4. A variational solution for a cylinder of finite length \n In the second stage of our argument, we attempt to include the inhomogeneity in the static \ndemagnetizing field by adopting the trial form \n \n± p m ( ) = Z(z) J (k r)cos(p ) ⊥ r . (2.31) 17 If we substitute this form in Eq. (2.17), together with some radially averaged z -dependent \nmagnetic field \nH(z) , we obtain the following one -dimensional eigenvalue problem \n \n22ex2dZ(z) H(z) D k Z(z) 0\ndx⊥ \n − − − = (2.32) \nwith the boundary conditions \n \nZ (0) = Z (h) = 0 . (2.33) \nIn the spirit of this variational approach we could obtain\nH(z) ) by averaging with respect to \n2pJ (k r) ⊥\n, which would lead to slight differences between modes with differing \nk⊥ . \nAlternatively, we can use the analytic expression for the dipole field along the cylinder axis \nzH (z,r 0) =\n that arises from spins which are fully aligned (as expected for the case where the \nexchange is totally dominant)51. Assuming a cylinder of radius a and height h, and setting z = 0 \nand z = h at the caps, the resulting demagnetization field along z -axis is \n \n( )\n( )demag 02 2 2 2h z zH r 0,z 2 M 2\nza h z a−= = − − − ++ −+ . (2.34) \n In the limit of \na / h 0→ , \n( ) demagH r 0,z 0 == (corresponding to an infinite rod) and in \nthe limit of \na / h→ , \n( ) demag 0H r 0,z 4 M = = − (corresponding to a thin disk). Fig. 2.2 shows \nthe resulting magnetic field for YIG cylinders having a diameter of 75nm and lengths of 75, 150, \n300, 600, and 1200nm as calculated from Eq. (2.34) (dashed lines) and along the r = 0 axis by \nOOMMF. The close correspondence arises from the dominance of exchange in these small \ndiameter samples. \n2.5. Zeros of mode functions and mode labels \n The demagnetizing field plays the role of an external potential in the Schrodinger equation \n(2.32), and the strong decrease in this field near the end caps leads to surface bound states or cap \nstates whose wav e functions die off exponentially away from the caps. In principle there could \nbe many bound states, but for our parameters we find only one state at each cap. All higher \nenergy states are extended along the z direction, and since the demagnetization field is \nessentially uniform in the bulk of the cylinder, their wavefunctions behave approximately as \nsinusoidal standing waves. 18 \n \n \n Let us now recall that for a one -dimensional Schrodinger equation with a reflection sym - \nmetric potential, the states with succ essively higher energy alternate in parity and have \nsuccessively increasing number of zeros, with the lowest energy state having no nodes and even \nparity. Further their wave functions must be mutually orthogonal. For our problem, these \ntheorems are satisfi ed as follows. The two cap states are nearly degenerate but they are admixed \nby tunneling to form even and odd parity states with zero and one node respectively. (See, \nhowever, the discussion in Sec. 7 on how the inclusion of dipole -dipole interactions mod ifies the \nenergy ordering.) The first extended state must then have even parity and two nodes. To be \napproximated by \n( ) z cos k (z h/2) − and to be orthogonal to the cap states, we must have \nzk 2 / h\n corresponding to \nz2 = and a negative end defect \nz . Higher extended states must \nhave higher values of \nz . In this way we see the need for the restriction \nz2 in Eq. (2.30) and \nfor the end d efect at the same time. \n For each value of p and \nrn , we could label the differently quantized modes along z by the \nnumber of their zeros. The lowest extended or bulk mode in any family with given p and \nrn Fig. 2.2. The dashed line shows the analytic demagnetization field calculated from Eq. \n(2.32) using \n=0 4 M 1750Oe for YIG cylinders with a diameter of 75nm and five different \nlengths of 75, 150, 300, 600, 1200nm i n a field of 2000Oe. For comparison the solid line \nshows the field computed by OOMMF along the line r = 0. Note how for long cylinders the \nfield profile near one cap is insensitive to the presence of the other cap. 19 would then have the label \nz2 = , while the cap mode would be labeled \nz0 = . This is \nunaestehetic and also does not differentiate between the physically different character of the cap \nmodes vis a vis the bulk mode s. We therefore label the bulk modes by an index \nzn , with \n \nzzn2= − (2.35) \nFor the cap modes we replace the number \nzn by the letters ‘g’ (gerade, even parity) and ‘u’ \n(ungerade , odd parity). For reference we summarize the correspondence between the number of \nzeros and the mode labels as follows: \n No. of zeros Mode label \n 0 g \n 1 u \n \nz \nzzn2= − \nThe modes are labeled by the scheme \nrz(p n n ) with the letters ‘g’ or ‘u’ for cap modes in lieu \nof \nzn . In particular the mode nominally identified as the uniform FMR mode has the label (000) \n(but see the discus sion in Sec. 7). \n2.6. Numerical results for the variational approximation and comparison with simulations \n We now describe some results from the numerical integration of Eq. (2.32) together with \nthe position dependence of \n( ) demagH r 0,z = given by Eq. (2.34). Imposing the boundary \ncondition (2.33) at the faces then yields \nZ(z) together with the eigenvalues \n( ) zrn (p,n ) = , \nwhere for the general case \nzrn (p,n ) denotes the eigenvalue for given va lues of the azimuthal \nand radial mode numbers, p and \nrn . Given that we have neglected the transverse dipolar field in \nobtaining Eq. (2.32) we expect the resulting eigenvalues to be most accurate in the limit of large \nzn\n mode numbers, and particularly when both \nr p 0 and n 0= = (which corresponds to \n0= in \nEq. (1.3)). \n The dashed lines in figure 2.3 show the resulting form of Z(z) for the lowest lying cap \nmode with p = 0 an d no radial nodes for cylinders with a diameter of 75nm and heights of 75, \n150, 300, 600, and 1200 nm in a field of 2000G. Note the approximately exponential decay of \nthe amplitude as we proceed deep into the interior for the longer samples confirming thei r \nsurface like character. Also shown are the OOMMF simulations obtained using procedures to be \noutlined below (the fact that their amplitudes do not go strictly to zero in longer samples arises 20 from a contamination from other modes). Accompanying anti symmetric modes (not shown) are \nalso highly localized while in addition having a node at the cylinder midpoint . \n As is evident, the semi -analytic results for Z(z) are surprisingly good for \nh 300nm . The \nfrequencies however are not. These c ould be improved by including the transverse dipolar field \nusing perturbation theory, which will raise the frequency. We have not attempted this exercise \nsince our approximate treatment is quite rough in the first place and it would not add to our \nqualitat ive understanding. \n \n \nFig. 2.3. The dashed lines show the behavior of the mode function Z(z) vs. z obtained from integrating Eq. \n(2.32) for the (00g) cap mode for cylinders with a diameter of 75nm and heights of 75, 150, 300, 600, and \n1200nm. The solid li nes show the OOMMF simulation results for the same parameters. \n \n As a crude estimate of the cap mode frequency we can compare it with the frequency of a \nhat box (disc) with a radius equal to its height. Reported demagnetization coefficients52 for this \naspec t ratio are \nN 0.4745=\n and \nN 0.2628⊥= . For a field \n0H 2.000 kOe= and \n0 4 M 1.750 kOe = \n Eq. (1.2) yields f = 4.568 GHz. For our exchange dominated sample it is \nreasonable to add a correction of order \n2\nexD / a 1.0 kOe = which raises the frequency to 5.6 \nGHz which is to be compared with the OOMMF value of 6.64 GHz. 21 \n \n Fig. 2.4. The dashed line shows the mode function Z(z) vs. z obtained from \nintegrating Eq. (2.35) for the (00g), (000), (002), a nd (004) modes. The solid line \nshows the corresponding forms arising from OOMMF. Note the behavior at the \ncylinder faces closely conforms with the boundary condition (2.24). \n \n Table I. End defect for (00n) modes. \nMode Label z hkz,fit/ z \n000 2 1.62 −0.38 \n001 3 2.50 −0.50 \n002 4 3.68 −0.32 \n003 5 4.78 −0.22 \n004 6 5.84 −0.16 \n005 7 6.96 −0.04 \n006 8 7.90 −0.10 \n007 9 8.92 −0.08 \n008 10 10.01 +0.01 \n009 11 10.91 −0.09 \n00,11 13 12.96 −0.04 \n00,13 15 14.98 −0.02 \n00,15 17 16.92 −0.08 \n Calculations for the extended states with mode number \nzn 0,1,= were also performed. \nHere we encounter progressively higher mode frequencies scaling approximately as \n2\nzn . Figure \n2.4 below shows the result of such calculations for the (00g), (000), (002), and (004) modes. \nTable I lists values of \nzk h/π , \nz and \nz for these and neighboring modes. Note that \nz0 → 22 with increasing \nz . We will explain why modes (00,10), (00,12), and (00,14) are not in this table \nat the end of Sec. 3.2. \n 3. Computational Approaches \n The material studied here is YIG which was chosen for its long mode lifetim es. Whether \nthese long lifetimes survive in submicron structures is an open question. The majority of the \nstudies were for a sample with h = 4d = 300 nm. The material parameters used are typical for \nYIG 53: γ = 2π × 2.8 GHz/kOe, saturation magnetization M s = 139emu/cm3, damping constant \n–5 =5 10\n, and exchange constant \n–7exA = 3.5 10 erg/cm. The applied field was 2 kOe \nalong z direction. Damping was turned on to relax the system to its initial state, and turned off \nafter the sy stem was excited for most simulations. In the few that it was not, it was too small to \nhave any significant effect. \n As noted above our simulations were carried out with the OOMMF code developed by the \nU. S. National Institute for Standards and Technology. This program divides a chosen sample \ninto cells on a rectilinear grid and numerically integrates the LL equation in time for their \nmagnetizations, \ni(t)M , as they evolve under the influence of the torques acting on them arising \nfrom a n external field, the nearest neighbor exchange interaction, and the dipolar fields of the \nremaining cells (anisotropy fields can be included but will be ignored in what follows). Each \nmagnetic moment is located at the center of the cell. The number of cel ls scales with the cube of \na characteristic sample dimension but was nominally fixed at 1/54 cells/(nm)3 corresponding to a \ncell size of \n3 nm 3 nm 6 nm for the \nd 75 nm, h 300 nm= = sample. There are 50 cells in the z \ndirection, and 489 c ells in the xy plane (489/625 = 0.7824 vs. /4 = 0.7854). In Sec. 7, we \nsimulate samples with other values of h. As described there, we then use cells with the same x \nand y dimensions (3 nm × 3 nm), but depending on the value of h, the dimension z is adjusted \nappropriately. \n3.1. Static equilibrium \n Prior to exciting the system, the spins were initially aligned along the cylinder axis \n(parallel to the external field) after which the system was evolved in time (with damping) until it \nstabilizes in an equilibrium configuration. Various tests can be ap plied to determine that it is a \nglobal equilibrium state. This part of our simulations yields the static magnetic field distribution \nwhich could also be used for the calculations in Section 2. 23 3.2. Exciting the system \n Several different excitation schemes were utilized. In the simplest of these, all spins were \ntipped by a small fixed angle relative to their equilibrium orientations in a plane containing the z \naxis as an initial condition. This favors the excitation of uniformly precessing modes. To drive a \nparticular non-uniform mode the spins were tipped from their equilibrium positions in a manner \nthat mimics the mode (such as that obtained as the mode pattern in a prior simulation)54. To drive \na broader spectrum of modes that is localized around a time \n0t and some position \n0r we tip the \nspins in some direction according to the function41 \n \ny0 0 x 0 x 0\n0 0 0 0sin[ k (y y )] sin[ (t t )] sin[ k (x x )] sin[ k (z z )]F(t, ) A(t t ) (x x ) (y y ) (z z )− − − −=− − − −r (3.1) \nwhere \nx y x, k , k and k control the extent to which the excitation is local ized in time and \nspace. Here x, y, z denote cell coordinates. Such pulses can also be introduced at multiple times \nand positions to favor the excitation of modes with differing spatial and temporal properties. In \nparticular inclusion of only the last facto r induces modes propagating along z. Forms can be \nconstructed that favor the excitation of radial or azimuthal modes. Finally, some simulations \nwere performed in which individual spins were tipped in random directions within some \nspecified average angular range. This excites a very broad range of modes and if the tipping \nangles are large (e.g., approaching 180o) generates a “hot” system, from which it is difficult to \nextract clear modes. Altogether, we tried more than ten different excitation pulses in an e ffort to \nidentify modes with different symmetry and numbers of nodes. Despite this, our mode table (see \nSec. 4) has gaps. In some cases, modes are nearly degenerate [for example modes (003), ( -10g), \nand ( -10u)] and cannot be easily resolved. In others, the y were too high in frequency to be seen \nwith the particular excitation pulse employed. We are confident that these modes exist and that \nour mode classification is complete. \n3.3. Identifying modes \n As the system state simulated by OOMMF evolves in time from some chosen initial \nconfiguration, the magnetization vectors \ni( ,t)mr at the (discrete) cell sites \nir are recorded at \nregular time intervals. From this data set we can perform a cell by cell fast -Fourier transfo rm \n(FFT) within some chosen time interval available from the simulation to obtain the complex \nquantities \ni( , ) mr . We stress that the OOMMF simulation does not assume that \neq()Mr is 24 along \nˆz or that the deviations \ni( ,t)mr are in the x -y plane, although the most useful information \nis contained in these components for low amplitude mode studies. From the FFT, we follow \nMcMichael and Stiles5 and construct the cell -wise power s pectra, \n \n2x i x iS ( , ) | m ( , ) | = rr , (3.2) \ntogether with their sum over the entire sample, \n \nx x i\niS ( ) S ( , ) = r , (3.3) \nand likewise for \nyiS ( , ) r and \nyS ( ) . As noted by them, this definition of a power spectrum is \nvery different from the power spectrum of the integrated magnetization (total magnetic moment \nof the sample), which is what makes them so useful in mode identification; in particular, the \nfrequencies where these total sample power sp ectra have sharp maxima are identified as possible \nmode frequencies of the system. As an example, the power spectra in Eq. (3.3) are given in Fig. \n3.1 which shows various modes including a very high frequency mode that is aliased to less than \n5 GHz due to a Nyquist critical frequency of 50 GHz . \n Suppose a mode has been identified at a frequency \na . With the sign conventions used in \nour numerical FFT program, the mode pattern associated with this frequency is given by \n \n( )a (a) i( t )ii( ,t) ( , )e += m r m r . (3.4) \nThe phase \n is arbitrary and amounts to a choice of the zero of time. To avoid unnecessary \nminus signs, we choose \n3 / 2 = which gives \n \n(a)i i a( ,0) ( , ) = m r m r , \n(a)i i a( ,T / 4) ( , ) = m r m r , (3.5) \nwith \na T 2 /= being the time period of the mode. Hence, by examining the imaginary and real \nparts of the vector \nia( , ) mr we can, respectively, obtain the real -space vector magnetization at \nsome ti me and a quarter cycle later for the spatial pattern associated with some nominal mode at \nthe specific frequency. By plotting these vector fields, we can get a highly visual depiction of the \nmode, permitting easy mode assignment and further analysis. \n 25 \n \nFig. 3.1. An example of an FFT spectrum. The left and right y axes show \n()2\nx sample2SωN and \n()2\nsampleSω2Nz\n. See Eq. (3.3). A broad sinc pulse as described in Eq. (3.1) was used to excite this \nspectrum . \n \nIn constructing the power spec trum, modes with higher frequencies than the inverse of the \nchosen integration time step, which violate the Nyquist sampling criterion and are then aliased to \nlower frequencies, must be identified and rejected. In the exchange dominated samples \nconsidered here some of them can be identified as spurious peaks with frequencies lower than \nthe known uniform modes, but in dipole dominated larger samples genuine modes below the \nuniform modes are expected. More generally they must be identified by altering the tim e interval \nover which the transform is performed to determine if some mode moves its position. Most \nsimulations were done with a time step of 10 ps and for a duration of 10.24 ns. \n A curious spatial aliasing was also observed (as evidenced by a rapid spati al variation of \nthe mode intensity on the scale of the cell period) in some patterns; it is thought to be associated \nwith a spatial FFT that is performed to calculate the dipole field in the underlying program. Such \nmodes must also be rejected. \n3.4. Impli cations of the p degeneracy for OOMMF patterns \n Using the procedures described we can construct mode maps in chosen planes by plotting \nthe complex cell amplitudes, \nxy ˆ ˆ ( , ) m ( , ) m ( , ) = + m r r x r y , at frequencies where the power 26 spectrum shows maxima. On the b asis of these patterns we are typically able to assign \napproximate mode numbers p, n r and n z, and designate them as \nr z r z(pn n ) (pn n )( , ) mr for that \nfrequency, \nrz(pn n ) = . The p mode number requires special attention as we now discuss. In \nwhat immediately follows we will drop the mode designation, regarding it as being understood. \n Writing the complex function \n( , ) mr in component form as \n \nx x x\ny y ym ( , ) m ( , ) im ( , )\n( , )m ( , ) m ( , ) im ( , ) + = = + r r r\nmrr r r , (3.6) \nthe corresponding behavior in the tim e domain follows as \n \nx x x it\ny y ym ( ,t) m ( , ) im ( , )\n( ,t) em ( ,t) m ( , ) im ( , )− + = = + r r r\nmrr r r \n \nxx\nyym ( , )cos t m ( , )sin t\nm ( , )cos t m ( , )sin t + = + rr\nrr . (3.7) \nIf the modes of the system have a pure p character, as in Eq. (2.11), we can write the above \ncomponents as \n \nxm ( , ) m(r,z,p, )cos(p ) = r , \nxm ( , ) m(r,z,p, )sin(p ) = r , (3.8a, b) \n \nym ( , ) m(r,z,p, )sin(p ) = − r , \nym ( , ) m(r,z,p, )cos(p ) = r . (3.8c, d) \nHence, we can write \n( ,t)mr as \n \nx\nym (r,z,p,t) cos(p t)( ,t) m(r,z,p, )m (r,z,p,t) sin(p t) − = = − − mr . (3.9) \nHere the magnetization vector rotates as \n changes with a radially symmetric amplitude. In our \napproximation, where the variables r and z separate, we would write the solutions that have even \nparity as \n \n( )x\n0 z pym ( , t) cos(p t)m cos k (z h / 2) J (k r)m ( , t) sin(p t)⊥ − =− − − r\nr . (3.10) \n If the modes of the system have a pure p character and in addition p and \np− are \ndegenerate we can form symmetric and anti symmetric standing wave super -positions of the two \nforms of Eq. (3.10) to obtain 27 \nxx\nyxm (r,z,p, t) m (r,z, p, t)\n( , t)m (r,z,p, t) m (r,z, p, t)− =−mr \n \ncos(p t) cos( p t) m(r,z,p, )sin(p t) sin( p t) − − − = − − − − \n (3.11) \nwhich results in the following two forms \n \ncos( t)m(r,z,p, )cos(p )sin( t) (3.12a) \n \nsin( t)m(r,z,p, )sin(p )cos( t) − , (3.12b) \nor if our model product form is assumed, \n \n( ) 0 z pcos( t)m cos k (z h / 2) J (k r)cos(p )sin( t)⊥ − , (3.13a) \n \n( ) 0 z psin( t)m cos k (z h / 2) J (k r)sin(p )cos( t)⊥ − − . (3.13b) \nHere again the spin direction rotates (in both senses) but now the amplitude is modulated in \n . \nA similar discussion applies to the solutions that are odd parity. \n In Appendix B we discuss how a pure p mode pattern can be e xtracted from a \nsuperposition of +p and –p OOMMF patterns. \n3.5. Separating nearly degenerate modes \n As noted above, in the absence of the dipole interaction modes with +p and –p are \ndegenerate. When this interaction is present such modes are split; i.e ., according to the \ndiscussion of section 3. 4 our eigen modes will be running waves in \n . But the splitting rapidly \ndecreases for larger mode numbers, and for running times less than \n1t− , where \n is the \nsplitting, the power spectrum displays a single (slightly broadened) peak at the mode frequency. \nIn order to resolve the splitting in a power spectrum the OOMMF run times must be increased. \n When the splitting is not resolved in the power spectrum the resulting mode patterns \ndisplay a standing wave character. For a few of these we used the standing wave mode pattern as \nan initial configuration and ran the program long enough to display a beat pattern in time from \nwhich the splitting c ould be accurately determined. An example of this technique is shown in \nFig. 3.2 for the cylinder with d = 75 nm and h = 300 nm. Fig. 3.2a shows the case of the \n( 105) 28 mode s with an average frequency of 17.29 GHz. Note a beat waist occ urs at t = 32.68 ns from \nwhich we calculate the mode splitting as 30. 60 MHz . \n \n \nFig. 3.2. Beat pattern s emerging from the time evolution arising from the splitting \nof (a) the \n( 105) modes with an average frequency 17.29 GHz and a spli tting of \n30.60 MHz , (b) the \n,r p 0 n 0== , g and u cap modes with an average frequency \n6.54 GHz and a splitting of 14.4MHz. \n 29 A very small splitting also occurs between the symmetric (or gerade, denoted g) and anti-\nsymmetric (ungerade, denot ed u) combinations of the cap modes. In the Schrodinger equation \nlanguage of Sec. 2, this is a tunnel splitting between the surface bound states. This splitting is \nintrinsically small and hard to resolve in long cylinders, although we have resolved it for the \nr p 0, n 0==\n, g and u cap modes for the cylinder with h = 300 nm. Now \nf 6.54 GHz= and \nf 14.4 MHz=\n. The corresponding beat pattern is shown in Fig. 3.2b. \n4. Frequencies of low -lying modes \n Most computation s were carried out on a YIG sample with \nh = 4d =8a = 300 nm in a \nstatic field of \n0H 2000 Oe= . To test the behavior at small and large \nzk some calculations were \ncarried out for samples with h = 7.5, 37.5, 75, 150, 600 and 1200 nm. The material parameters \nused are typical for YIG, as given earlier in section 3. \n Table II lists the frequencies \nr z r zpn n pn nf / 2 = of low lying modes as obtained from \nthe peaks in the power spectrum; all entries are for \nh = 4d =8a = 300 nm and in a static field of \n0H 2000 Oe=\n. The mode numbers come from a comparison of the accompanying mode pattern \nwith the forms discussed in Sec. 2 with special attention to the number of radial and longitudinal \nzeros and how \nm winds around the z axis. By fitting to the form (2.21) we can assign discrete \nwavevectors, \nr,nk⊥ and \nznk ; these values are used for a comparison with the HK formula as we \ndescribe in the next section. All modes with \np0 have a node at \nr = 0 ; for larger values of \nk⊥ , \nadditional radial nodes can be present and their mode number is denoted as \nrn0 . Modes listed \nas \np are nearly degenerate in the sense discussed above, and the mode patterns display \nstanding waves in \n as described by Eq. (3.9a,b). \nTable II. The mode frequencies for a YIG cylinder with a height of 300nm and a diameter of \n75nm in a static magnetic field \n0H = 2000 Oe, organized into families with given mode \nnumbers p and n r. Multiply listed modes (e.g., ( –115) and (\n 115)) we re observed with both \npure and with mixed +p and –p character, depending on the methodology used to extract \nthem (e.g. FFT vs. beat pattern) . \n \n 30 \n \n Table II \nMode Frequencies vs. p, \n rzn , and n \np nr nz f (GHz) p nr nz f (GHz) p nr nz f (GHz ) \n0 0 Even cap (g) 6.543 1 0 g 10.35 ±1 0 5 17.29 \n0 0 Odd cap (u) 6.543 1 0 u 10.35 ±1 0 7 21.78 \n0 0 0 7.813 1 0 1 11.91 ±1 0 11 33.98 \n0 0 1 8.105 1 0 3 13.96 ±1 1 g 35.16 \n0 0 2 8.887 1 0 5 17.29 ±1 1 u 35.16 \n0 0 3 10.06 1 0 7 21.78 ±1 1 1 36.91 \n0 0 4 11.52 1 0 9 27.44 ±1 1 3 39.26 \n0 0 5 13.38 1 0 11 33.98 ±1 1 5 42.68 \n0 0 6 15.53 1 0 13 41.5 ±1 1 9 52.83 \n0 0 7 17.97 0 1 0 22.66 ±2 0 g 15.82 \n0 0 8 20.7 0 1 1 23.24 ±2 0 u 15.82 \n0 0 9 23.63 0 1 2 24.22 ±2 0 1 18.07 \n0 0 11 30.27 0 1 3 25.49 ±2 0 2 18.36 \n0 0 13 37.79 0 1 4 27.05 ±2 0 3 19.63 \n0 0 15 46.09 0 1 5 28.91 ±2 0 5 22.95 \n−1 0 g 10.06 0 1 6 30.96 ±2 0 9 33.11 \n−1 0 u 10.06 0 1 7 33.4 ±2 0 11 39.65 \n−1 0 1 11.72 0 1 8 36.04 ±2 0 15 55.37 \n−1 0 2 12.7 0 1 9 38.96 ±3 0 g 24.02 \n−1 0 3 13.87 0 1 10 42.19 ±3 0 u 24.02 \n−1 0 4 15.43 0 1 11 45.51 ±3 0 1 25.59 \n−1 0 5 17.19 0 1 12 49.12 ±3 0 3 27.83 \n−1 0 6 19.34 −1 1 g 35.16 ±3 0 5 31.25 \n−1 0 7 21.78 −1 1 u 35.16 ±3 0 7 35.84 \n−1 0 8 24.41 −1 1 3 39.26 ±3 0 11 47.95 \n−1 0 9 27.34 −1 1 5 42.68 \n−1 0 10 30.57 −1 1 6 44.82 \n−1 0 11 33.98 0 2 g 55.57 \n−1 0 12 37.6 0 2 0 56.64 \n−1 0 13 41.5 0 2 2 58.3 \n−1 0 14 45.51 0 2 4 61.23 \n−1 0 15 49.71 0 2 6 65.53 \n 0 2 8 70.41 31 \n \n \n \n \nFig. 5.1. a) Typical fit of the radial OOMMF amplitude to the function \n1J (k r) ⊥ for \nthe ( –105) mode. b) T ypical fit of the axial OOMMF amplitude to the function \n()z cosk z h / 2 −\n for the (006) mode. \n \n 32 5. Comparison of simulated frequencies with the Herring -Kittel expression \n It is interesting to examine the extent to which the OOMMF mode frequencies in Table II \ncan be represented by the Herring -Kittel frequencies as given by Eq, (1.3). To do this we need \nvalues of \nzk and \nrp,nk for the extended modes. A preliminary value of \nzk follows from \ncounting the number of nodes along z. Better values emerge55 from fitting \nrm (z,r 0, ) = to \n( ) z cos k (z h / 2) −\n or \n( ) z sin k (z h / 2) − . Values for \np,nk were obtained by fitting \nrm (z,r, ) to \np p,nJ (k r)\n at some z with \np,nk as an adjustable parameter. Fig. 5.1a, b shows examples of such \nfits. Note that although we do not employ it to find \nzk and k⊥ , the boundary condition at the \nfaces for this mode closely approximate s maximum amplitude as opposed to maximum \nderivative. \n We now show some plots of the frequencies, \nr z r z(pn n ) (pn n )f / 2 = inferred from the \nsimulations, for various modes \nrz(pn n ) versus \nzk at fixed \np,nk with these latter values \nobtained by the above procedures. Also shown are the frequencies predicted by the H - K \nexpression, Eq. (1.3), for the same wave -vector components and a demagnetization coefficient of \nN 0.098=\n. \n The \n symbols in Fig. 5.2 show the results of the OOMMF simulation for the \nz (00n )\n modes, as a function of \nzk in units of \n/h , for which the lowest frequency is \n7.81GHz. Not included are the accompanying cap -mode frequencies which are both 6.41GHz \nwithin the resolution. The continuous curve shows the frequencies predicted by the Herring -\nKittel (H - K) formula. The ag reement is surprisingly good, especially at small \nzk where H - K \nis expected to break down. To some extent this may arise from the fact that the internal magnetic \nfield is position dependent, being lower at the caps, and thereby pr oducing an effect partially \ncompensating that of \nN⊥ . The \n symbols show the predictions of the one -dimensional \nSchrodinger equation as described in section 2.2. \n \n \n \n 33 \n \nFig. 5.2. The frequencies of the \nz (00 n ) modes vs. \nzk as obtained from OOMMF \nsimulations ( symbols), Herring -Kittel formula (continuous line) and the \nSchrodinger equation (\n ), as discussed in section 2.2. For the Schrodi nger \nequation data \nzk h / is a proxy for \nzn2+ . \n \n \nFig. 5.3. The symbols show the OOMMF simulations for the mode frequencies versus the \ndimensionless wave -vector \nzk for p = 0 and \nr n 0, 1 and 2= ; the continuous curves show \nthe predictions of the H – K expression. 34 \n The symbols in Fig. 5.3 show the OOMMF simulations of p = 0 mode frequencies as a function \nof the dimensionless \nzk for \nrn 0, 1 and 2= while the continuous curves show the predictions for the \ncorresponding mode numbers of the H - K formula. Again, the agreement is excellent. Readers may \nnotice that some modes are missing in these plots. This is because they were not excited with t he \nprotocols used, but we are confident they exist and that their mode patterns conform with the general \nframework presented here. \n \nFig. 5.4. The OOMMF mode frequencies vs. \nzk for \nrn0= and \n − p 0, 1, 2, 3=\n. The curves show the predictions of the H – K equation. \n \n \n Lastly, Fig. 5.4 shows the OOMMF simulations for mode frequencies vs. \nzk for \nrn0= \nand \np 0, 1, 2, 3 = − . The H - K formula again gives an excellent overall representation. \n Analogous to our approximating the position dependence of the bulk states with a form \nz cos(k z)\n we can use \nze− and \n(z h)e− to qualitatively describe the amplitude in the vicinity of \nthe cylinder faces for the cap mode; i.e., we take k as imaginary by writing \nki= , in which case \n2k\n is replaced by \n2− in a Herring -Kittel like expres sion, which pushes the frequency below \nthat of the first extended mode. \n5.1. Why the H - K formula works so well 35 The H - K relation is sometimes referred to as a spin -wave dispersion relation, and indeed \nit is tantamount to saying that the normal modes o f the body can be described by a continuous (or \nquasicontinuous) variable k. As explained in Appendix A, this assertion is grossly incorrect \nwhen the spatial variation of m is on a scale comparable to the dimensions of the body56; the \nmode functions must th en take account of the shape of the body, and be described by discrete \nsets of appropriate mode numbers, which may or may not be wavevector -like. \n Nevertheless, if it should turn out that some discrete modes with slow spatial variation are \nwell described b y wavevector -like variables, then for those modes the H - K expression can be \nexpected to give the frequency to good approximation, especially if one is in the exchange \ndominated limit. This is the situation in the present investigation. We have seen that for every \nbulk mode, we can identify a reasonably well -defined \nzk and an orbital angular momentum \nquantum number p. We also see a radial dependence in \n(r)m that pretty well matches a Bessel \nfunction \npJ (k r) ⊥ from which can obtain a \nk⊥ . \n That the mode functions should look this way is not an accident. We have some support for \nthis functional behavior from the theory of the infinite cylinder in the exchange dominated l imit. \nWe intend to publish the details of this theory separately, and here we only summarize the key \nresults. In first order perturbation in the small parameter \n \n0\n2exM= \nDk (5.1) \nwe find that the mode function is given by \n \n( )zx i(p t) i((p 2) t) ik zp\nym 11 1J (k r) f (r) e f (r)e em ii 2− − + + − ⊥ = + + − . \n (5.2) \nThe corrections \nf are both \nO( ) . We add here that we discovered the properties of the mode \nfunctions by experimenting with OOMMF first, and developed the theoret ical framework later. \n6. Examples of mode patterns \nTable II in section 4 lists approximately 90 modes which we have identified. We will \nnow present some accompanying mode patterns which display various behaviors. All figures in \nthis section pertain to YIG cylinders with a diameter d = 2a = 75 nm and a height h = 300 nm in \na magnetic field H 0 = 2000 Oe. 36 There is a wealth of information in these figures as we now explain. They all depict \nvarious aspects of the frequency space Fourier amplitude m(r, ), which is a complex vector, \ni.e., its x, y, and z components are all complex numbers. The real parts make a vector, and so do \nthe imaginary parts, which we can call the real and imaginary parts of the complex vector, \n( , )mr\n and \n( , )mr . We discard the z component leaving the xy projection m⊥(r, ). \nCircular panels such as Figs. 6.1a and 6.1b show xy cross sections of the cylinder, while \nrectangular panels such as Fig. 6.1c show yz cross sections through a diameter of the cylinder. In \nthe circular panels, the arrows show the directions and relative magnitudes of either \n( , )⊥mr \nor \n( , )⊥mr , while the thin black lines show contour levels of the magnitudes of these same \nvectors, i .e., either \n( , )⊥mr or \n( , )⊥mr . These contour levels are also color coded \naccording to the scale on the right. As explained in Sec. 3.3 — see Eqs. (3.4) and (3.5) — the \n \npanel shows the spi ns a quarter -cycle after the \n panel. That is, time proceeds from \n to \n , \nwhich is reflected in the sequence of panels a) and b) when both \n and \n parts are shown. In \nthe rectangular panels we show contours of \nxm ( , ) r ; again, the contours are color coded. \nBecause the circular panels show the magnitude of the vector in the xy plane, while the \nrectangular panels show only the x component, the contour levels in the two types of panels \ncannot be directly compared. Furthermore, depending on just how a particular mode is excited, \nthe spins can have larger projections along the x or y directions at the particular time captured in \nthe xy cross sections, and this can further affect the values of the contour levels in the rectangular \npanels vis-a-vis the circular ones. The most salient feature is the variation or the relative Fourier \namplitude within a panel. We add here that our simulations are done with 50 vertical layers of \ncells of height 6 nm each. There is a layer extending from z = 144 to 150 nm, and another from z \n= 150 to 156 nm. Hence, circular panels such as in Figs. 6.1a and 6.1b that are labeled z = 147 \nnm cor respond to the midpoint of the cell layer just below the midplane of the cylinder; panels \nlabeled z = 3 nm such as in Fig. 6.2a show the lowest layer; panels labeled z = 39 nm such as in \nFigs. 6.5a and 6.5b show the 7th layer from the bottom. However, in the text and figure captions \nwe have described the panels at z = 3 nm and z = 147 nm as lying at z = 0 and z = h/2, as this is \nmore natural and intuitively easier to understand. \nIn the interest of clarity, we shall repeat these points as necessary, and add further \ninformation about the patterns as we discuss them one by one. 37 We start with the lowest lying modes: the nominal bulk uniform precession or cylindrical \nKittel (0, 0, 0) mode with f = 7.813 GHz, which is concentrated within the body of the cylinder, \naway from the caps, together with the even (0, 0, g) and odd (0, 0, u) cap modes concentrated on \nthe top and bottom cylinder faces with a mean frequency of 6.543 GHz and a splitting that is too \nsmall to be resolved. \nFig. 6.1a shows \n()000(x,y,z h 2, )⊥ = m while Fig. 6.1b shows \n()000(x,y,z h 2, )⊥ = m\n for the (0, 0, 0) bulk mode with f = 7.813 GHz; here \n( , , 2, )x y z h⊥ = m\n and \n( , , 2, )x y z h⊥ = m denote the normalized vector fields of the \nFourier amplitude given by following prescript ions: \ni\n12\n2 1\ncell i\ni( , )K\nN ( , )⊥\n−\n\n \nmr\nmr\n and \ni\n12\n2 1\ncell i\ni( , )K\nN ( , )⊥\n−\n\n \nmr\nmr (6.1) \nwhere K is a global constant scale factor whose value is chosen as 1000 for convenience for all \nmode patterns (this and subsequent ones). The arrows indicate the xy projection of the \nmagnetization. The lines and color coding depict the contours of constant amplitude, either \n()i,⊥mr\n or \n()i,⊥mr . Ideally these would be concentric circles but there is always \ncontamination at some level from other mode s. There may also be numerical errors associated \nwith the discretization. Time proceeds from \n (panel (a)) to \n (panel (b)), a quarter cycle later. \nFig. 6.1c shows \n()000\nxm (x 0,y,z, ) = , again with contour lines together with color \ncoding. Note the contour lines are quite parallel to the faces, a behavior that arises from the \nstrong influence of exchange in these small samples and validates the factorized form Eq. (2.21) \nfor \n(), F r z− . \nFig. 6.2a shows \n()00g(x,y,z 0, )⊥ = m for the symmetric (g) cap mode with f = 6.543 \nGHz. Figs. 6.2b and 6.2c show \n()00g\nxm (x 0,y,z, ) = and \n()00u\nxm (x 0,y,z, ) = for the even \n(g) and odd (u) cap modes respectively. We see that the mod e intensity is strongly concentrated 38 near the cylinder faces, dropping off rapidly as one proceeds to the interior. Note the anti -\nsymmetric character of the u mode is clearly apparent as seen from the node at z = h/2. \nFigs. 6.3a, b show \n( )10u(x,y,z 0, )−\n⊥ = m and \n( )10u(x,y,z 0, )−\n⊥ = m of the (−1, 0, u) \nantisymmetric 10.06 GHz cap mode which has a node at r = 0. Note how the spins wind through \nan angle of 2 as we proceed counter -clockwise around the line r = 0. Fig. 6.3c shows \n( )10u\nxm (x 0,y,z, )− = \n where the antisymmetric behavior in z is evident. \nFigs. 6.4a, b show \n()10u(x,y,z 0, )⊥ = m and \n()10u(x,y,z 0, )⊥ = m for the \nneighboring p = +1 mode with a frequency of 10.35 GHz. Here one encounters the “retrograde” \nmotion associated with the oppositely winding sense of m with the azimuthal angle . \nWe next consider a mode with multiple nodes along z. As remarked earlier the splitting \nbetween ±p modes diminishes as the overall mode numbers increase, so we designate them with \nboth signs since our mode projection method generally yields a superpositi on. Here we consider \nthe (±1, 0, 5) mode(s) with f = 17.29 GHz. Figs. 6.5a, b show \n( )105(x,y,z 39nm, )\n⊥ = m and \n( )105(x,y,z 39nm, )\n⊥ = m\n while Fig. 6.5c shows \n( )105(x 0,y,z, )\n⊥ = m . (We recall that z = \n39 nm is the midplane of the 7th layer of cells from the bottom of the cylinder.) Note the mode \npatterns in the xy plane now display nodes since the modes with p = +1 and p = −1 interfere to \nform a partial standing wave. Our plot in the yz plane contains the two end nodes arising from \northogonality to the cap modes discussed above (the patterns for which we do not show) as well \nas the five interior nodes. If we use the projecti on technique described in Appendix B, we can \nagain separate the two modes. This is shown in Figs. 6.6a and 6.6b where we plot \n( )105(x,y,z 39nm, )+\n⊥ = m\n and \n( )105(x,y,z 39nm, )−\n⊥ = m . These are also the modes for \nwhich we resolved the splitting via the beat pattern in Fig. 3.2a. \nAs an example of a mode with a larger azimuthal mode number and mixed p character, \nFig. 6.7 shows a plot of \n( )305(x,y,z 39nm, )\n⊥ = m , which has a frequency of 31.25 GHz. 39 Finally, we present a mode with additional radial nodes. Such a mode will have a high frequency \nconsidering the relatively small diameter of our sample. Figs. 6.8a and 6.8b show \n()020(x,y,z h 2, )⊥ = m\n and \n()020\nxm (x 0,y,z, ) = for the (020) mode with a frequency of f \n= 56.64 GHz. \n \n 40 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Fig. 6.1. Mode pattern for the lowest (Kittel -like) bulk (0 0 0) mode, with a frequency of \n7.813 GHz . Parts a) and b) show an x - y cross section of th e imaginary and real \nparts of the Fourier transform amplitude through the cylinder mid point while part \nc) is the real y - z cross section cont aining the cylinder axis. The arrows show the \ndirection of the spins. The spin orientations for the real part correspond to a time 1/4 \ncycle later than that for the imaginary part. The lines show contours of constant \n| ( , ) | mr\n in (a) and (b) and of constant \n( , )x mr in (c) ; these values are also color \ncoded according to th e scale given to the right. \n 41 Fig. 6.2. The p = 0, n r = 0 cap \nmodes. Part a) shows the real x - y \ncross section at z = 0 of the \n(00g)\nsymmetric cap mode; this \nmode has the globally lowest \nfrequency of 6.543 GHz . Parts b) \nand c) show the real y - z cross \nsections of the g and u modes \ncontaining the cylinder axis of the \nfrom which the surface \nconfinement is apparent . \n \n 42 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Fig. 6.3. Parts a) and b) show the imaginary and real x - y cross \nsections at z = 0 of the ( –1 0 u) antisymmetric cap mode with a \nfrequency of 10.06 GHz. Part c) is the real y - z cross section \ncontaining the cylinder axis from which the surface \nconfinement is apparent. \n \n 43 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \nFig. 6.4. Parts a) and b) show th e imaginary and real x - y cross sections at z = 0 of the \n(10u) antisymmetric cap mode with a frequency of 10.35 GHz. Important to note here is \nthe retrograde character of the spin winding. 44 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Fig. 6.5. Parts a) and b) show the imaginary and real x - y cross sections at z = 39 nm of the \n( 105)\n modes with a frequency of 17.29 GHz. Note the standing wave behavior of these \ncross -sections arising from the superposition of azimuthally counter propagating mode s. \nHowever, at a fixed point in space the spins still precess counterclockwise. Part c) is the real y \n- z cross section containing the cylinder axis. \n \n 45 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n Fig. 6.6. Here the projection technique described in Appendix B has been used to separate the modes \nwith p = +1 and p = –1 from the same data used to construct figure 6.5. Note the azimuthal intensity is \nnow constant as appropriate for running waves. \n \n \nFig. 6.7. The real x - y cross section of \nthe \n( 305) bulk mode with a \nfrequency o f 31.25 GHz showing a \nmultiplicity of azimuthal nodes. Note \nalso the deep central node due to the \n3r\n behavior of \n3J (k r) ⊥ . \n 46 \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \nFig. 7.1. The depen dence of the frequency on h for the lowest \nzk bulk mode (000), symmetric cap \nmode (00g), and antisymmetric cap mode (00u) of a YIG cylinder with d =75nm. Also shown is the \nfrequency predicted by the Kittel expression. The inset sh ows the region where the symmetric cap \nmode crosses over the antisymmetric cap mode for small h. In the region below about 1 00nm the \n(00g) mode replaces the (000) mode as the quasi uniform mode we associate with the Kittel formula. \nSee discussion. \n Fig. 6.8. The real x -y and y -z cross -sections of the (02 0) mode with a frequency \nof 56.64 GHz exhibiting multiple radial nodes. \n \n 47 7. Evol ution of the low -lying mode behavior with cylinder height \n Although the majority of our simulations for mode patterns were carried out for a YIG \ncylinder with d = 2a = 75 nm and h = 300 nm, the behavior of the three lowest lying p = 0 \nmodes, (000), (00g), and (00u), was studied over the much wider range of h extending from \n7.5nm to 1200 nm57. Fig. 7.1 shows the frequencies emerging from the OOMMF simulations \ntogether with the predictions based on the Kittel expression, Eq. (1.2), according to the \ndemagnetiza tion coefficients found by Joseph and Schl omann58. Importantly we see that the \n(00g) “cap mode” evolves into the dominant mode in the thin -disc limit , while the (000) bulk \nmode becomes dominant (although lying slightly higher in frequency) in the long cyli nder limit. \nThe Kittel formula, which is a single equation, actually describes two different modes in these \nlimits and does not apply well to any mode for 150 nm \n h \n 300 nm . The level crossing of the \n(00g) and (00u) modes does not violate the Wigner -von Neumann anti -crossing theorem due to \ntheir even and odd character. Furthermore, the apparent violation of the ordering of energy levels \nof 1d Schrodinger equations is resolved by noting that the dipole -dipole interaction is a non-local \nperturbation which puts the eigenvalue problem outside the Sturm -Liouville class. \n \nFig. 7.2. The contours of constant intensity for three modes of a YIG cylinder with d = 75 nm and h = 37.5 \nnm. a) The 4.59 GHz symmetric (00 g) cap mode, corresponding to the Kittel mode in the thin -disc limit; note \nit has no nodes. b) The 13.57 GHz antisymmetric (00u) cap mode. c) The 42.97 GHz (000) “bulk” mode. \n \n It is clear from this discussion that the true behavior of the dominant FMR re sponse \ncannot be described by a simple Kittel -like expression; a characterization in terms of 48 demagnetization coefficients is often inadequate and glosses over the spatial complexity of the \nmode with the greatest spectral weight in a uniformly excited samp le. \n Some mode patterns for a YIG cylinder with d = 75 nm and h = 37.5 nm are shown in \nfigure 7.2. The cell size for these simulations is 3 x 3 x 1.5 nm^3 . Fig7.2a shows the shows the \n4.59 GHz symmetric (00g) “cap” mode which for these dimensions has becom e the dominant \nmode. It is a cap mode in name only since it spans the entire sample. Rather than the planar \nconstant amplitude contours encountered in the longer cylinders, this mode now has \napproximately cylindrical ones. Figure 7.2b shows the antisymmetric cap mode for a cylinder \nwith the same dimensions which now has a significantly higher frequency of 13.57 GHz; here \nthe contours of constant amplitude run approximately parallel to the faces indicating that the odd \nmode has substantially changed charact er from the even one. Lastly, we show the bulk mode in \nFig. 7.2c. Consistent with the results in Fig. 7.1 this mode has the highest frequency, 42.97 GHz. \nThe contours of constant amplitude run approximately parallel to the faces and we obtain a quite \nregul ar sine wave with one complete wavelength along z. \n \n8. Conclusions and possible applications \n We have explored the mode structure of nanoscopic cylinders of yttrium iron garnet, both \nanalytically and through many -spin simulations, in a regime where the ef fects of exchange \ndominate the response. In addition to the extended (bulk) modes, which can be classified in \nterms of the azimuthal, radial, and axial mode numbers designated as \nrzp,n ,n , we find \nsymmetric and antisymmetric combination s of cap modes, that are localized at each of the \ncylinder faces. In all cases they lie lower in frequency than the accompanying family of modes \nzn\n for given azimuthal and radial mode numbers \nr p and n . When exami ning the height \ndependence, we find that the dominant FMR response cannot be precisely described by a simple \nKittel -like formula. How this picture would change in passing to the dipole dominated limit is \ndeserving of additional study. \n By way of applicatio ns, there is a growing effort directed at using magnetic bits for \ncomputation as well as data storage. In particular it has been demonstrated that the required logic \noperations can be accomplished with lines and arrays of dipole -coupled, single -domain, bar \nmagnets with dimensions of a few hundred nanometers59. While promising, this approach is still 49 restricted to what can be done using binary macro -spin flips (and cascades thereof) of the \nindividual magnets. \n Looking farther ahead it is natural to ask if lo gic functions can be performed by \nexploiting the internal dynamics of nano particles . For the case of wave -guide based operations \nthis is already a world -wide activity60. But here we envision exciting (and mixing) large \namplitude resonant modes within a sin gle nanoparticl e. This can involve a single or multiple \ninputs applied simultaneously or sequentially, with different microwave frequencies and/or \npolarizations . When the particles are small the various modes are well separated and can be \naddressed individ ually and rapidly. By exploiting intrinsic nonlinearities and optimizing the \nsample dimensions (to tune the mode frequencies), different pump frequencies can be efficiently \nmixed, a topic we are exploring independently. Progress in this direction requires an \nunderstanding of the low amplitude mode structure of the particles involved. \n Apart from the cap modes, the modes studied here are all extended in character and in \nthat sense are the standing wave counter part of the plane wave modes that were the start ing \npoint of the non -linear analysis carried out by Suhl discussed in the introduction . However \nongoing simulations of nano -scale cylinders and elliptical discs at large precession amplitudes \ndisplay instabilities involving the edge -nucleation of dynamic vortex and antivortex modes. \nPossibly related instabiliti es occur a t domain walls in stripes61,62. The connection between such \nstates and low -lying extended mode s in understanding large amplitude precession dynamics in \nnanomagnets is currently unclear. \n \nAcknowledgement \n This research was carried out under the support of U.S. Department of Energy through \ngrant DE -SC0014424. This research was supported in part through the computational resources \nand staff contributions provided for the Quest high performance computing facility at \nNorthwestern University which is jointly supported by the Office of the Provost, the Office for \nResearch, and Northwestern University Information Technology. \n \nAppendix A: The Herring - Kittel equation \n The Herring - Kittel (H - K) exp ression, given earlier as Eq. (1.3), is 50 \n2 2 2 2 2in,z ex in,z 0 ex(H D k )(H 4 M sin D k ) = + + + (A.1) \nwhere we wrote \nin,z 0 0H H 4 N M − \n with \nN\n an axial demagnetization coefficient, and \n \nexex\n02ADM= , (A.2) \nwhich fixes the exchange energy density used in the OOMMF simulation, \n \njj exex2iii,j0MM AExx M= . (A.3) \nTo understand the remarkable agreement between our OOMMF results and this formula, \nlet us recall how it is derived. \n The linearized Landau -Lifshitz eq uation is \n \nex 2in,z 0\n0d 2AˆHMdt M= − + mz m h m (A.4) \nwhere h is the dynamic demagnetization (or dipolar) field induced by m. Fourier transforming \nwith respect to space and assuming a time dependence, \nite− we obtain \n \nex 2k in,z k 0 k k\n02Aˆ i H M kM− = − −m z m h m . (A.5) \nThe field \nh is governed by the Maxwell equations, \n \n = 4π , = 0 − h m h , (A.6) \ntogether with the requirements that the normal component of \n4 = + b h m and the tangential \ncompon ent of h be continuous at the boundary of the particle. If we now Fourier transform Eq.’s \n(A.6), and simply ignore the effects of the boundary conditions we obtain \n \nkk24\nk= − kmhk . (A.7) \nInserting (A.7) into Eq. (A.5), requiring th e equations for the resulting two components to be \ncompatible, and taking \nin,zH to be homogeneous, leads immediately to the HK formula, Eq. \n(A.1). \n From the above derivation we see that, qualitatively, the HK approximation accounts f or \nthe axial (static) demagnetization but neglects some part of the transverse contribution. Since the \ndipole -dipole interaction is long ranged, this neglect is qualitatively profound, and quantitatively 51 valid only for short wavelengths, when \nka 1 . When this condition is satisfied we can argue \nthat the magnetic charges induced on the “lateral” surface by m change sign rapidly on a length \nscale \n1k− , so the field produced by them dies off on the same length scal e and may be ignored in \nthe bulk of the particle. To further clarify this behavior, we will derive the HK formula in a \nsecond way. \n As is known from magnetostatics, the (normalized) magnetic field, \n()hr , can be regarded \nas arising f rom a magnetic charge density \n()mr according to \n \n3\n3 V()( ) 4 ( )d r\n||+− = − \n−rrh r m r\nrr . (A.8) \nThe integral here is taken to extend infinitesimally beyond the particle volume, as indicated by \nthe superscript in \nV+ . In this way both volume and surface charges are included. If we Fourier \ntransform this equation, we obtain \n \n3\n24 d k \nk− = − \nkk k kkmhk (A.9) \nwhere \nk is the Fourier transform of unity over the particle volume: \n \ni3\n3V1e d r\n(2 )+−=\nkrk . (A.10) \nWe can write (A.9) more compactly as \n \nk24\nk= − kkkmhk (A.11) \nwhere ∗ denotes a convolution. For a cylinder of height h and radius a, \n \n2z1\n3za h sin k h 2J (k a)k h k a (2 )⊥\n⊥=\nk . (A.12) \nWhen \nzk h 1\n and \nka⊥ \n in the convolution in Eq. (A.11), \nk can be approximat ed by a \ndelta -function, \n()k , and we recover Eq. (A.7). For smaller k this approximation is invalid. \n \nAppendix B. Relation between standing and running waves in \n \nThe magnetization fields corresponding to running waves associated with orbital angular \nmomentum p and –p (with p > 0) have the form, 52 \n()() ( ) ( ) p ˆ ˆ ,t F r,z cos p t sin p t= − − − m r x y , (B.1a) \n \n()() ( ) ( ) p ˆ ˆ ,t F r,z cos p t sin p t− = − − − − − m r x y , (B.1b) \nwhere\n F(r,z) is an unspecified function. By superposing these fiel ds, we obtain a standing wave \npattern, \n \n() ()() () ()s\np p p ˆ ˆ ,t F r,z cos p cos t sin t− = + = + m r m m x y . (B.2) \nWe wish to recover the running wave patterns from a knowledge of the standing wave pattern. \nTo do this, we first transform the latter by rotating the amplitude by p, and the ve ctor by \ndirection by . The transformed field is, \n \n() ( )( ) ( )\n()() () ()Ts\npp,t r,z, 2p ,t 2\nˆ ˆ F r,z sin p sin t cos t .= − + \n= − + m r m\nxy (B.3) \nIt is now easy to see that the difference of the transformed and the original stationary wave \npattern gives us \npm : \n \n() () ()sT\np p p,t ,t ,t 2 =−m r m r m r . (B.4a) \nLikewise, the sum gives \np−m : \n \n() () ()sT\np p p,t ,t ,t 2−=+m r m r m r . (B.4b) \n What this means is the following. Supposing OOMMF has produced a pattern which has \np nodal lines in at some fixed time. We denote this patte rn by \ns\npm as above. We then consider \nthe vector fields \n \n( )\n( )( )\n( )( )\n( )ss\np,x p,y p,x\nssp,yp,y p,xm r,z, m r,z, 2p m r,z, 11\n22 m r,z, m r,z, m r,z, 2p\n − − = − \n . (B.5) \nIf these combinations are used as initial conditions in OOMMF, they should evolve into \np+ and \np−\n running waves as indicated. In this way we can obtain positive confirmation that these \nrunning waves are indeed eigenmodes; this is how we separated the \np+ and \np− modes shown in \nFig.’s 6.5 and 6.6. \n \n References 53 \n1 J. H. E Griffiths, Anomalous High -frequency Resistance of Ferromagnetic Metals , Nature, 158, \n670 (1946), DOI: 10.1038/158670a0. \n2 H. Suhl, Subsidary Absorption Peaks in Ferromagnetic Resonance at High Signal Levels, Phys. \nRev. 101, 1437 (1956). \n3 H. Suhl, The Theory of Ferromagnetic Resonance as High Signal Powers, J. Phys. Chem. Of \nSolids 1, 209 -227 (1957). \n4 M. J. Donahue and D. G. Porter, OOMMF User's Guide, Version 1.0, NISTIR 6376 , \nNational Institute of Standards and Technology, Gaithersburg, MD (Sept 1999). \n5 R. D. McMichael and M. D. Stiles, Magnetic Normal Modes of Nanoelements, J. Appld. Phys. \n97, 10590 (2005); DOI: 10.1063/1.1852191. \n6 Jian-Gang Zhu, Xiaochun Zhu, and Yuhui Tang, Micro wave assisted magnetic recording, IEEE \nTrans. Magn. 44, 125 -131 (2007). \n7 K. Rivkin, M. Benakli, N. Tabat, and H. Yin, Physical principles of microwave assisted \nmagnetic recording , J. Appl. Phys. 115, 214312 (2014). \n8 S. Okamoto, N. Kikuchi, M. Furuta, O. Kitakami and T. Shimatsu, Microwave assisted \nmagnetic recording technologies and related physics J. Phys. D: Appl. Phys. 48, 3530011 \n(2015). \n9 K. Rivkin, N. Tabat, and S. Foss -Schroeder, Time -dependent fields and anisotropy dominated \nmagnetic media , Appl. Phys. Lett. 92, 153104 (2008) . \n10 Jinho Lim, Zhaohui Zhang, Anupam Garg, and John Ketterson, Simulating Resonant \nMagnetization Reversals in Nanomagnets, IEEE Trans. Magn. 57, 1300304 (2021) DOI: \n10.1109/TMAG.2020.3039468. \n11 Jinho Lim, Zhaohui Zhang, Anupam Garg and John B. Ketterson, Pi Pulses in a Ferromagnet: \nSimulations for Yttrium Iron Garnet, J. Mag. Mag. Mater. 527, 167787 (2021). DOI: \nhttps://doi.org/10.1016/j.jmmm.2021.167787 \n12 Eiichi Fukushima and Stephan B. W. Roeder , Experimental Pulsed NMR: A Nut s and Bolts \nApproach, CRC Press, 1981. \n13 Christophe Thirion, Wolfgang Wernsdorfer, and Dominique Mailly, Switching of \nmagnetization by nonlinear resonance studied in single nanoparticles, Nature Materials, Vol 2, \nPg. 524 (2003); DOI:10.1038/nmat946 . \n14 H. Sato, T. Kano, T. Nagasawa, K. Mizushima, and R. Sato, Magnetization switching of a \nCo/Pt multilayered perpendicular nanomagnet assisted by a microwave field with time varying \nfrequency, Phys. Rev. Applied 9, 054011 (2018). \n15 Y. Li, et. al., Nutation spec troscopy of a nanomagnet driven into deeply nonlinear \nferromagnetic resonance, Phys. Rev. X 9, 041036 (2019). \n16 For a recent example of zero -field resonances in a nano -scale object see: Wonbae Bang, F. \nMontoncello, M. T. Kaffash, A. Hoffmann, J. B. Ketters on, and M. B. Jungfleisch, \nFerromagnetic resonance spectra of permalloy nano -ellipses as building blocks for complex \nmagnonic lattices, J. Appl. Phys. 126, 203902 (2019); https://doi.org/10.1063/1.5126679. \n17 If technical issues associated with growing high quality YIG films on non -magnetic substrates \ncan be solved (i.e., not GGG), achieving long mode lifetimes at low temperatures is \nanticipated. Nanocylinders can then be patterned and excited by superconducting antennas \nwherein close coupling can be achieve d together with locally tailored field profiles to more \neffectively couple to specific modes. 54 \n18 A. Yamaguchi, K. Motoi, A. Hirohata, H. Miyajima, Y. Miyashita, and Y. Sanada, Broadband \nferromagnetic resonance of Ni 81Fe19 wires using a rectifying effect, Phys. Rev. B 78, 104401 \n(2008), DOI: 10.1103/PhysRevB.78.104401 . \n19 U. Ebels, J. L. Duvail, P. E. Wigen, L. Piraux, L. D. Buda, and K. Ounadjela, Ferromagnetic \nresonance studies of Ni nanowire arrays, Phys . Rev. B 64, 144421 (2001), DOI: \n10.1103/PhysRevB.64.1 44421. \n20 C. A. Ramos, M. Vazquez, K. Nielsch, K. Pirota, J. Rivas, R. B. Wehrspohn, M. Tovar, R. D. \nSanchez and U. Gösele, J. Mag. Mag. Mater. 272–276, Part 3, 1652 (2004), \ndoi.org/10.1016/j.jmmm.2003.12.233 . \n21 A del Campo and C Greiner, SU -8: A photores ist for high -aspect -ratio and 3D submicron \nlithography, J. Micromech. Microeng. 17, R81 (2007) doi:10.1088/0960 -1317/17/6/R01. \n22 L. D. Landau and E. M. Lifshitz, Course in Theoretical Physics, Statistical Physics Part II \nPergamon Press, Oxford, 1980, Secti on 69. \n23 C. Kittel, On the Theory of Ferromagnetic Resonance Absorption, Phys. Rev. 73 (2): 155 –161 \n(1948), doi:10.1103/PhysRev.73.155. \n24 Exciting such modes with a uniform microwave field requires they have an even number of \nnodes, otherwise the net inter action averages to zero. To address this problem, highly efficient \nantennas having multiple elements that are spaced so as to spatially resonate with some set of \nmodes have recently been employed: Jinho Lim, Wonbae Bang, Jonathan Trossman, Andreas \nKreise, Matthias Benjamin Jungfleisch, Axel Hoffmann, C. C. Tsai, and John B. Ketterson, \nDirect detection of multiple backward volume modes in yttrium iron garnet at micron scale \nwavelengths, Phys. Rev. B 99, 014435 (2019). \n25 F. Bloch, On the theory of ferromagnet ism, Z. Physik 61, 206 (1930); On the Theory of the \n Exchange Problem and the Appearence of Retentive Ferromagnetic , Z. Physik 74, 295 (1932). \n26 A. M. Clogston, H. Suhl, L. R. Walker and P. W. Anderson , Possible Source of Line Width in \nFerromagnetic Reso nance, Phys. Rev. 101, 903 (1956). \n27 A. M. Clogston, H. Suhl, L. R. Walker and P. W. Anderson , Ferromagnetic Line Width in \nInsulating Materials, J. Phys. Chem. Solids 1, 129 (1956). \n28 J. E. Mercereau and R. P. Feynman, Physical Conditions for Ferromagnetic Resonance, Phys. \nRev. 104, 63 (1956). \n29 L. R. Walker, Magnetostatic Modes in Ferromagnetic Resonance, Phys. Rev. 105, 390 (1957). \n30 R. W. Damon and J. R. Eshbach, Magnetostatic modes of a ferromagnet slab, J. Phys. Chem. \nSolids 19, 308 (1961), doi.org/10. 1016/0022 -3697(61)90041 -5. \n31 R. W. Damon and H. Van De Vaart, Propagation of Magnetostatic Spin Waves at Microwave \nFrequencies in a Normally ‐Magnetized Disk, J. Appl. Phys. 36, 3453 (1965); \nhttps://doi.org/10.1063/1.1703018 \n32 D. E. De Wames and T. Wolfram, Dipole -exchange spin waves in ferromagnetic films, J. \nAppl. Phys. 41, 987 (1970); doi.org/10.1063/1.1659049. \n33 R. E. Arias, Spin -wave modes of ferromagnetic films, Phys. Rev. B 94, 134408 (2016). \n34 R. I. Joseph, and E. Schlomann, Theory of magnetostati c modes in long, axially magnetized \ncylinders, J. Appl. Phys . 32, 1001 (1961). \n35 R. Arias and D. L. Mills, Theory of spin excitations and the microwave response of cylindrical \nferromagnetic nanowires, Phys. Rev. B 63, 134439 (2001). \n36 C. Herring and C. Ki ttel, On the theory of spin waves in ferromagnetic media, Phys. Rev. 81, \n869 (1951) . 55 \n37 S. Shultz and G. Dunifer, Observation of spin waves in sodium and potassium, Phys. Rev. \nLett. 18, 283 (1967). \n38 N. Masuhara, D. Candela, D. O. Edwards, R. F. Hoyt, H. N. Scholz, D. S. Sherrill and R. \nCombescot, Collisionless Spin Waves in Liquid 3He, Phys. Rev. Lett. 53 (12), 1168 (1984). \n39 V. P. Silin, Oscillations of a Fermi -liquid in a magnetic field, Zh. Eksp. Teor. Fiz. 33, 1227 \n(1957) [Sov. Phys. JETP 6, 945 (1958)] . \n40 A. J. Leggett, Spin diffusion and spin echoes in liquid 3He at low temperature, J. Phys. C : \nSolid State Phys. 3, 448 (1970) . \n41 G. Venkat, D.Kumar, M. Franchin, O. Dmytriiev, M. Mruczkiewicz, H. Fangohr, A.Barman, \nM. Krawczyk, and A. Prabhakar, Proposal for a Standard Micromagnetic Problem: Spin Wave \nDispersion in a Magnonic Waveguide IEEE Trans. Magn . 49, 524 –529 (2013). \n42 L. D. Landau and E. M. Lifshitz, Statistical Physics Part II, Pergamon Press, Section 69; the \nform written here follows Herring and Kittel. \n43 This argument is standard, and can be found in many books. See, e.g., R. Shankar, Principles \nof Quantum Mechanics, Plenum, New York, 1980, Exercise 12.5.1 and Fig. 12.1. \n44 See, e.g., Anupam Garg, Classical Electromagnetism in a Nutshell, Princeto n University Press, \nPrinceton, N. J., 2012, pp. 174 –177, and 239 –240. \n45 See Eqs. (19), (23), and (24) in Ref. 34. \n46 See the unnumbered equation two equations above Eq. (22) in Ref. 29. \n47 H. A. Bethe and E. E. Salpeter, Quantum Mechanics of One and Two Elec tron Atoms, \nSpringer, Berlin (1957), Sections 29 and 30. \n48 W. S. Ament and G. T. Rado, Electromagnetic Effects of Spin Wave Resonance in Ferro -\nmagnetic Metals, Phys. Rev. 97, 1558 (1955) \n49 Amikam Aharoni, Introduction to the Theory of Ferromagnetism, Oxfor d, 1996, p. 178. \n50 See, e.g., C. M. Bender and S. Orszag, Advanced Mathematical Methods for Scientists and \nEngineers, McGraw -Hill, 1978, Chapter 9. \n51 This has the virtue of replacing the discrete OOMMF field profile with a continuous one; it \nalso gave some what better agreement with the simulated OOMMF mode profiles. \n52 M. Sato and Y. Ishii, Simple and approximate expressions of demagnetizing factors of \nuniformly magnetized rectangular rod and cylinder, J. Appl. Phys. 66 (2), 983 (1989). DOI: \n10.1063/1.343481 . \n53 A. G. Gurevich and G. A. Melkov, Magnetization oscillations and waves, CRC Press, Boca \nRaton, 1996. \n54 It can also drive additional modes having some overlap on the chosen mode at some initial \ninstant in time. \n55 Alternatively, we could Fourier transform the simulated form of \nrm (z,r, ) along z, where the \nbar indicates a radial average, and identify the peak in the spectrum. Since our description is \nonly semi quantitative this procedure was not attempted. \n56 In particular Eq. (A.11) cannot be replaced by Eq. (A.7) in Appendix A. In this regard, we find \nthe discussion by L. R. Walker, Resonant Modes of Ferromagnetic Spheroids, J. Appl. Phys. \n29, 318 -324 (1958) and his Fig. 6 to be particularly germane. \n57 The simulations were done with cell d imensions of 3 nm × 3nm in the x and y directions, and \nthe dimension z along the z direction was adjusted so as to give varying numbers of layers \ndepending on the height h. The number of layers varied from 10 for h = 7.5 nm, to 25 for h= \n16.8 to 145 nm, a nd either 25 or 50 for h ≥ 150nm. The choice of 25 layers is enough to 56 \ncapture the spatial variation of the three modes we are seeking in this section. It is more \nimportant to increase the run time (by as much as a factor of 8) in order to resolve the mode s in \nfrequency. \n58 R. I. Joseph and E. Schlomann, Demagnetizing field in non -ellipsoidal bodies, J. Appl. Phys. \n36, 1579 (1965). \n59 For a review see: M. T. Niemier, G. H. Bernstein, G. Csaba, A. Dingler, X. S. Hu, S. Kurtz, S. \nLiu, J. Nahas, W. Porod, M. Siddiq and E. Varga, Nanomagnet logic: progress toward system -\nlevel integration, J. Phys. Condens. Matter 23, 493202 (2011) . \n60 Abdulqader Mahmoud, Florin Ciubotaru, Frederic Vanderveken, Andrii V. Chumak, Said \nHamdioui, Christoph Adelmann, and Sorin Cotofana, Introduction to spin wave computing, J. \nAppl. Phys. 128, 161101 (2020); https://doi.org/10.1063/5.0019328 \n61 D. J. Clarke, O. A. Tretiakov, G. -W. Chern, Ya. B. Bazaliy, and O. Tchernyshyov, Dynamics \nof a vortex domain wall in a magne tic nanostrip: Application of the collective -coordinate \napproach, Phys. Rev. B 78, 134412 (2008). \n62 M. Klaui, Topical Review: Head -to-head domain walls in magnetic nanostructures, J. Phys. \nCondens. Matter 20, 313001 (2008). " }, { "title": "2102.02116v2.Infinite_Series_of_Ferrimagnetic_Phases_Emergent_from_the_Gapless_Spin_Liquid_Phase_of_Mixed_Diamond_Chains.pdf", "content": "arXiv:2102.02116v2 [cond-mat.str-el] 6 Apr 2021Journal of the Physical Society of Japan FULL PAPERS\nInfinite Series of Ferrimagnetic Phases Emergent from the Ga pless Spin\nLiquid Phase of Mixed Diamond Chains\nKazuo Hida∗\nProfessor Emeritus, Division of Material Science, Graduat e School of Science and Engineering,\nSaitama University, Saitama, Saitama, 338-8570\n(Received )\nThe ground-state phases of mixed diamond chains with ( S,τ(1),τ(2)) = (1/2,1/2,1), where Sis the\nmagnitude of vertex spins, and τ(1)andτ(2)are those of apical spins, are investigated. The apical spin s\nτ(1)andτ(2)are connected with each other by an exchange coupling λ. Other exchange couplings are set\nequal to unity. This model has an infinite number of local cons ervation laws. For large λ, the ground state\nis equivalent to that of the uniform spin 1 /2 chain. Hence, the ground state is a gapless spin liquid. For\nλ≤0, the ground state is a Lieb-Mattis ferrimagnetic phase wit h spontaneous magnetization msp= 1\nper unit cell. For intermediate λ, we find a series of ferrimagnetic phases with msp= 1/pwhereptakes\npositive integer values. The phases with p≥2 are accompanied by the spontaneous breakdown of the\np-fold translational symmetry. It is suggested that the phas e with arbitrarily large p, namely infinitesimal\nspontaneous magnetization, is allowed as λapproaches the transition point to the gapless spin liquid p hase.\n1. Introduction\nThe quantum effects in low-dimensional frustrated\nmagnets have been extensively studied in recent con-\ndensed matter physics.1,2)Various exotic quantum\nphases emerge from the interplay of quantum fluctuation\nand frustration. Among them, the diamond chain,3–9)\nwhose lattice structure is shown in Fig. 1, is known as a\nmodel with an infinite number of local conservationlaws.\nThe ground states can be classified by the correspond-\ning quantum numbers. If the two apical spins have equal\nmagnitudes, which is the case widely investigated, each\npair of apical spins can form a singlet dimer. It cuts the\ncorrelation between both sides and the ground state is a\ndirect product of the cluster ground states separated by\ndimers.\nThe ground states of spin-1/2 diamond chains have\nbeen investigated in Ref. 3. In addition to the spin clus-\nter ground states, the ferrimagnetic state with sponta-\nneous magnetization msp= 1/2 per unit cell is found.\nIn the latter phase, the apical spins form triplet dimers\nand all the spins collectively form a long-range ordered\nferrimagnetic state. Extensive experimental studies have\nbeen also carried out on the magnetic properties of nat-\nural mineral azurite that is regarded as an example of\ndistorted spin-1/2 diamond chains.8,9)\nThe ground states of spin-1 diamond chains have been\nalso investigated in Refs. 3 and 4. In addition to the spin\ncluster ground states, the nonmagnetic Haldane state\nand the ferrimagnetic states with spontaneous magne-\ntizationmsp= 1 and 1/2 are found. It should be noted\nthat the latter ferrimagnetic state is accompanied by a\nspontaneous translational symmetry breakdown.\n∗E-mail address: hida@mail.saitama-u.ac.jpOn the other hand, if the magnitudes of the two apical\nspins are not equal to each other, they cannot form a sin-\nglet dimer. Hence, all spins in the chain inevitably form a\nmany-body correlatedstate. In manycases, (quasi-)long-\nrange order evolves including the vertex spins. As a sim-\nple example of such cases, we investigate the mixed di-\namond chain with apical spins of magnitude 1 and 1/2,\nand vertex spins, 1/2 in the present work. Remarkably,\nwe find an infinite series of ferrimagnetic phases.\nThispaperisorganizedasfollows.InSect.2,themodel\nHamiltonian is presented. In Sect. 3, the ground-state\nphase diagram is determined numerically. The behavior\nof the spontaneous magnetization in each phase is pre-\nsented and analyzed. The last section is devoted to a\nsummary and discussion.\n2. Hamiltonian\nWe consider the Hamiltonian\nH=L/summationdisplay\nl=1/bracketleftBig\nSl(τ(1)\nl+τ(2)\nl)\n+(τ(1)\nl+τ(2)\nl)Sl+1+λτ(1)\nlτ(2)\nl/bracketrightBig\n,(1)\nwhereSl,τ(1)\nlandτ(2)\nlare spin operators with magni-\ntudesSl=τ(1)\nl= 1/2 andτ(2)\nl= 1. The number of the\nunit cells is denoted by L, and the total number of sites\nis 3L. Here, the parameter λcontrols the frustration as\ndepicted in Fig. 1.\nThe Hamiltonian (1) has a series of local conservation\nlaws. To see it, we rewrite Eq. (1) in the form,\nH=L/summationdisplay\nl=1/bracketleftbigg\nSlTl+TlSl+1+λ\n2/parenleftbigg\nT2\nl−11\n4/parenrightbigg/bracketrightbigg\n,(2)\n1J. Phys. Soc. Jpn. FULL PAPERS\nSlλ\n11τl(1)\nSl+1\nτl(2)11 S=τ(1)=1/2\nτ(2)=1\nFig. 1. Structure of the diamond chain investigated in this work\nS=τ(1)= 1/2 andτ(2)= 1.\nwhere the composite spin operators Tlare defined as\nTl≡τ(1)\nl+τ(2)\nl(l= 1,2,···,L).(3)\nThen, it is evident that\n[T2\nl,H] = 0 (l= 1,2,···,L). (4)\nThus, we have Lconserved quantities T2\nlfor alll. By\ndefining the magnitude Tlof the composite spin Tlby\nT2\nl=Tl(Tl+1), we have a set of good quantum numbers\n{Tl;l= 1,2,...L}whereTl= 1/2 and 3/2. The total\nHilbertspaceoftheHamiltonian(2)consistsofseparated\nsubspaces, each of which is specified by a definite set of\n{Tl},i.e.,asequenceof1/2and3/2.Apairofapicalspins\nwithTl= 1/2 is called a doublet (hereafter abbreviated\nas d) and that with Tl= 3/2 a quartet (abbreviated as\nq).\n3. Ground-State Phase Diagram\n3.1 Ground states for λ≫1\nForλ≫1,∀l Tl= 1/2. Hence, this model is equiva-\nlent to the spin-1/2 antiferromagnetic Heisenberg chain\nwhose ground state is a gapless spin liquid.\n3.2 Ground states for λ≪1\nForλ≪1,∀l Tl= 3/2. Hence, this model is equiv-\nalent to the spin-1/2-3/2 alternating antiferromagnetic\nHeisenberg chain whose ground state is a ferrimagnetic\nstate with spontaneous magnetization msp= 1 per unit\ncell according to the Lieb-Mattis theorem.10)Here,msp\nis defined by\nmsp=1\nLL/summationdisplay\nl=1(/an}bracketle{tSz\nl/an}bracketri}ht+/an}bracketle{tTz\nl/an}bracketri}ht) (5)\nwhere/an}bracketle{t/an}bracketri}htdenotes the expectation value in the ground\nstate with an infinitesimal symmetry breaking magnetic\nfield inz-direction.\n3.3 Intermediate λ\nIn the absence of spontaneous translational symmetry\nbreakdown, only above two phases are allowed. To pur-\nsue the possibility of other phases, we employ the finite\nsize DMRG method with the geometry of Fig. 2. TheSlλ\n11τl(1)\nSl+1\nτl(2)11\nFig. 2. Lattice structure used for the finite size DMRG calcula-\ntion\ncorresponding Hamiltonian is given by\nH=L/summationdisplay\nl=1SlTl+L−1/summationdisplay\nl=1TlSl+1+L/summationdisplay\nl=1λ\n2/parenleftbigg\nT2\nl−11\n4/parenrightbigg\n.(6)\nThis geometry is chosen to allow for the nonmagnetic\nground state for λ≫1. Here and in what follows, the\nnumber of states χkept in each subsystem in the DMRG\ncalculation ranged from 240 to 360. We find that the\nresults with χ= 240 are accurate enough in the present\nwork.\nThe ground-state energies of the Hamiltonian (6) up\ntoL= 16 for all possible configurations {Tl}are calcu-\nlated. The configurations that give the lowest energy are\nidentified for each λ. The spontaneous magnetizationper\nunit cell is given by\nmsp=1\nL/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingleL/summationdisplay\nl=1(Tl−Sl)/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle(7)\nfromtheLieb-Mattistheorem10)fortheHamiltonian(6).\nFigure 3 shows our numerical result for the λ-\ndependence of msp. In addition to the two phases de-\nscribed in Sect. 3.1 and Sect 3.2, a quantized ferrimag-\nnetic ground state with msp= 1/2 is clearly observed.\nThis state has the configuration with Tl= q for odd\nlandTl= d for even l. We denote this configuration\nas (qd)L/2. In the thermodynamic limit, this configura-\ntion tends to the (qd)∞configuration.This configuration\nis equivalent to the (dq)∞configuration in the thermo-\ndynamic limit. Hence, this phase is doubly degenerate\nand is accompanied by the spontaneous breakdown of\ntwofold translational symmetry. In what follows, similar\nnotations are employed for other configurations. Within\nthe finite-size DMRG calculation, the intermediate fer-\nrimagnetic phases are observed between this phase and\nthe nonmagnetic phase. Although two intermediate fer-\nrimagnetic phases are also observed near msp= 1, these\nareartifactsofthefinite sizeeffect, sincethey correspond\ntomsp= 1−1/Land 1−2/L.\nFor finite chains, however, the values of mspare con-\nstrainedbythe systemsize.Hence, itis notclearwhether\nmspis quantized or continuously varying with λfor\n0< msp<1/2 in the thermodynamic limit.\nHence, we resort to the infinite-size DMRG calcula-\ntion. However, it is not possible to carry out the calcu-\nlation for all possible configurations of {Tl}for infinite\n2J. Phys. Soc. Jpn. FULL PAPERS\n0.4 0.6 0.800.51\nλmsp\nL=8\nL=10\nL=12\nL=14L=14\nL=16(qd)∞\nFig. 3. λ-dependence of mspcalculated by the finite size DMRG\nmethod.\nS S S\nsingle q and (p−1) d’s\np S’s (S=1/2)1 segment = p unit cells\n= qdp−1q d d q\nFig. 4. (qdp−1)∞configuration.\nchains. Hence, we analyze this regime in the following\nway: We start with plausible candidates of periodic con-\nfigurations of Tl’s and find the configuration that gives\nthe lowest energy ground state among them. This leads\nto a stepwise λ-dependence of msp. Then, we check the\nstability of these steps against the formation of defects.\nAs plausible candidates of the ground states, we consider\nthe configurations(qdp−1)∞withmsp= 1/pthat consist\nof an infinite array of segments qdp−1with length of p\nunit cells as depicted in Fig. 4. These configurations can\nbe obtained from the configuration d∞corresponding to\nthe nonmagnetic ground state by inserting a q every p\nunit cells periodically. It should be remarked that these\nstates are accompanied by the p-fold spontaneous trans-\nlational symmetry breakdown. Hereafter, this series of\nconfigurations is called the main series configurations.\nThe case p= 1 corresponds to the q∞configuration that\ncorresponds to the Lieb-Mattis type ferrimagnetic phase\nwithmsp= 1.\nTheλ-dependence of mspin the main series is shown\nin Fig. 5. The spontaneous magnetization msprises con-\ntinuously from msp= 0 at the critical value of λgiven0.6 0.7 0.800.51\nλmsp\np≤38λc(∞)−~0.807p=2p=1\np=3\np=4\nFig. 5. λ-dependence of mspfor the main series configurations\ncalculated by the infinite size DMRG method.\n0 0.05 0.1 0.150.780.790.80.81λ\n1/p\nFig. 6. Extrapolation scheme of λc(p,p−1) andλc(p+1,p) to\np→ ∞. The filled and open circles correspond to the fits by Eq.\n(9) and Eq. (10), respectively. The filled square is the extra polated\nvalueλc(∞).\nby\nλc(∞)≡lim\np→∞λc(p,p−1)≃0.807 (8)\nwhere the boundary between the (qdp−1)∞phase with\nmsp= 1/pand (qdp−2)∞phase with msp= 1/(p−1) is\ndenoted by λc(p,p−1). The extrapolation to p→ ∞is\ncarried out using the data for 38 ≥p≥11 assuming the\nfollowing two asymptotic forms\nλc(p,p−1) =λc(∞)+C1\np+C2\np2, (9)\n3J. Phys. Soc. Jpn. FULL PAPERS\n0.8 0.802 0.804 0.806 0.80800.020.04\nλmsp\np≤38λc(∞)−~0.807\nEqs. (9),(10)\nFig. 7. λ-dependence of mspfor the main series configurations\nnearλ=λc(∞). The open circles are the results of the infinite\nsize DMRG calculation. The small filled circles and solid lin es are\nextrapolation by Eqs. (9) and (10).\nqdp−1qdp−1qdp−2qdp−1qdp−2\nn segments=qdp−2+(n−1)×qdp−1(a)\nqdp−1qdp−1qdpqdp−1qdp\nn segments=qdp+(n−1)×qdp−1(b)\nFig. 8. (a) (qdp−2(qdp−1)n−1)∞configuration and (b)\nqdp((qdp−1)n−1)∞configuration.\nλc(p+1,p) =λc(∞)+C′\n1\np+C′\n2\np2.(10)\nThe extrapolation procedure is plotted in Fig. 6. The\nboth extrapolations give the same value for λc(∞) up\nto the above digit. It should be noted that Eq. (9) and\nEq. (10) correspond to the extrapolation of left and right\nends of the steps, respectively. Using Eqs. (9) and (10),\ntheλ-dependence of mspis plotted down to msp= 0 in\nFig. 7.\nThe remaining question is whether the intermediate\nconfigurations with 1 /p < m sp<1/(p−1) can be a\nground state. To answer this question, it is necessary to\ncalculatethe ground-stateenergiesforall possibleconfig-\nurations of {Tl}, which is impossible. Hence, we confine\nourselves to the following configurations that are plausi-\nble to compete with the (qdp−1)∞configurations.\n(1) (qdp−2(qdp−1)n−1)∞configuration( msp=n/(p(n−1) + (p−1))) depicted in Fig. 8(a): A qdp−1seg-\nment is replaced by a qdp−2segment per every n\nsegments in the (qdp−1)∞configuration. For n= 1,\nthis configuration reduces to the (qdp−2)∞configu-\nration with msp= 1/(p−1) that corresponds to the\nneighboring step of the main series with higher msp.\nForn= 2, this configuration reduces to that with\nalternating qdp−1and qdp−2segments.\n(2) (qdp(qdp−1)n−1)∞configuration ( msp=\nn/(p(n−1)+(p+1))) depicted in Fig. 8(b):\nA qdp−1segment is replaced by a qdpsegment per\neverynsegments in the (qdp−1)∞configuration.\nForn= 1, this configuration reduces to the\n(qdp)∞configuration with msp= 1/(p+1) that\ncorresponds to the neighboring step of the main\nseries with lower msp. Forn= 2, this configuration\nreduces to that with alternating qdp−1and qdp\nsegments.\n(3) The configurations with spatial periodicity less than\n12 and spontaneous magnetization msp≤1/4 have\nbeen thoroughly checked.\nFor configurations (1) and (2), the computational cost\nincreases with an increase of pandn. We have numeri-\ncally confirmed that these states with 2 ≤n≤4 are not\nthe ground state for 2 ≤p≤11. It is further confirmed\nthat those with 2 ≤n≤6 are not the ground state for\n2≤p≤8. Within our numerical accuracy, we find no\nconfigurations that give lower energy than the main se-\nries configurations. Within the available numerical data,\nthe configurations with higher nare even less favorable.\nHence, it is highly plausible that the state intermedi-\natemspis not a ground state at least for p≤11. Also,\nconfigurations (3) do not give the ground state except\nfor the main series configurations. Thus, we expect that\nonly the configurations in the main series are realized in\nthe ground state.\n4. Summary and Discussion\nThe ground-state phases of diamond chains (1) with\n(S,τ(1),τ(2)) = (1/2,1/2,1) are investigated. Between\nthe gapless spin liquid phase for large λand Lieb-Mattis\nferrimagnetic phase with msp= 1 forλ≤0, we find a\nseries of quantized ferrimagnetic phases with msp= 1/p\nwhereptakes all positive integer values.\nTheλ-dependence of the spontaneous magnetization\nmspis very different from other diamond chains with fer-\nrimagnetic ground states. Although the quantized ferri-\nmagneticphasesarepresenteveninundistorteddiamond\nchains, the allowed values of spontaneous magnetization\nare limited to several simple rational values.3,4)\nThere are some examples of ferrimagnetic ground\nstates of diamond chains that are induced by the lat-\ntice distortion.5,6)In these cases, the ground state of\nthe undistorted chain is a paramagnetic state consist-\ning of clusters with finite magnetic moments. The sizes\n4J. Phys. Soc. Jpn. FULL PAPERS\nof the clusters are limited even in the absence of dis-\ntortion. The distortions induce ferromagnetic interac-\ntions between the cluster spins leading to the quantized\nferrimagnetic phases. The quantum fluctuations of the\nlengths of clusters are also induced by distortion leading\nto the partial ferrimagnetic phases.5)\nIn the present case, the ferrimagnetic phases are\npresent even in the absence of distortion. In contrast to\nthe cases of Refs. 3–6, arbitrarily large segments are al-\nlowed leading to the infinitesimally small steps around\nλ=λc(∞). However, no fluctuations of the lengths of\nthe segments are allowed in the present model, since the\nmagnitudes of the composite spins Tlremain good quan-\ntum numbers. Hence, partial ferrimagnetic phases are\nabsent in contrast to the cases of Refs. 5 and 6.\nIn the spin-1 alternating bond diamond chain with\nbond alternation δ, the nonmagnetic phase is equivalent\ntothegroundstateofthespin-1alternatingbondHeisen-\nberg chain with bond alternation δ.7)In this model, an\nintermediate ferrimagnetic phase is observed in the tiny\nregion close neighborhood of the point ( λ,δ) = (λc,δc)≃\n(1.0832,0.2598) that corresponds to the endpoint of the\nHaldane-dimer critical line.11–14)In Ref. 7, it has been\nspeculated that this region is the partial ferrimagnetic\nphase. However, considering the similarity of the gapless\nspin-liquid phase of the present model and the Haldane-\ndimercriticallineofthespin-1alternatingbonddiamond\nchain, it would be more reasonable to speculate that the\ninfinite series of quantized ferrimagnetic phases similar\nto those discussed in the present work is realized also in\nthis case. Unfortunately, the numerical confirmation is\ndifficult due to the smallness of the width of this region.\nAs mentioned above, in some examples of the quan-\ntized ferrimagnetic phases in undistorted diamond\nchains, the allowed values of spontaneous magnetization\nare limited to several rational values.3,4)In these exam-\nples, the nonmagnetic phases neighboring the ferrimag-\nnetic phases are spin-gap phases. On the other hand, the\nnonmagnetic phase neighboring the ferrimagnetic phase\nin the present model with infinitesimal step is the gap-\nless spin liquid phase. This seems to suggest that the\ninfinitesimal energy scale of the gapless spin liquid phase\nhelps the emergence of the exotic ferrimagnetic phase\nwith infinitesimal spontaneous magnetization. A further\nanalytical approach would be required to get insight into\nthe physical implication of the present phenomenon.So far, the infinite series of ferrimagnetic phases pro-\nposed in this work have not been found in real mate-\nrials. However, since the gapless spin liquid phases are\ngeneric critical states in quantum spin chains, it would\nbe possible that these series of phases are realized in\nthe presence of appropriate frustrating exchange inter-\nactions. Nevertheless, in more realistic cases, the pertur-\nbation that does not preserve the conservation laws (4)\nis inevitable. In such cases, the infinitesimal structure\nof spontaneous magnetization might be smeared. In this\ncontext, it would be an interesting problem to investi-\ngate the effect of lattice distortion in the present model.\nThese studies are left for future investigation.\nA part of the numerical computation in this work has\nbeencarriedoutusingthe facilitiesofthe Supercomputer\nCenter, Institute for Solid State Physics, University of\nTokyo,andYukawaInstituteComputerFacilityatKyoto\nUniversity.\n1)Introduction to Frustrated Magnetism: Materials, Experi-\nments, Theory , ed. C. Lacroix, P. Mendels, and F. Mila\n(Springer Series in Solid-State Sciences, Springer, Heide lberg,\n2011).\n2)Frustrated Spin Systems , ed. H. T. Diep, (World Scientific,\nSingapore, 2013) 2nd ed.\n3) K. Takano, K. Kubo, and H. Sakamoto, J. Phys.: Condens.\nMatter8, 6405 (1996).\n4) K. Hida and K. Takano, J. Phys. Soc. Jpn. 86, 033707 (2017).\n5) K. Hida, K. Takano, and H. Suzuki, J. Phys. Soc. Jpn. 79,\n114703 (2010).\n6) K. Hida, J. Phys. Soc. Jpn. 88, 074705 (2019).\n7) K. Hida, J. Phys. Soc. Jpn. 89, 024709 (2020).\n8) H. Kikuchi, Y. Fujii, M. Chiba, S. Mitsudo, T. Idehara, T.\nTonegawa, K. Okamoto, T. Sakai, T. Kuwai, and H. Ohta,\nPhys. Rev. Lett. 94, 227201 (2005).\n9) H.Kikuchi,Y.Fujii,M.Chiba,S.Mitsudo,T.Idehara,T.T one-\ngawa, K. Okamoto, T. Sakai, T. Kuwai T, K. Kindo, A. Mat-\nsuo, W.Higemoto, K.Nishiyama, M.Horovi´ c, and C.Bertheir ,\nProg. Theor. Phys. Suppl. 159, 1 (2005).\n10) E. Lieb and D. Mattis, J. Math. Phys. 3, 749 (1962).\n11) Y. Kato and A. Tanaka, J. Phys. Soc. Jpn. 63, 1277 (1994).\n12) S. Yamamoto, J. Phys. Soc. Jpn. 63, 4327 (1994); Phys. Rev.\nB51, 16128 (1995).\n13) K.Totsuka, Y.Nishiyama, N.Hatano, and M.Suzuki, J.Phy s.:\nCondens. Matter 7, 4895 (1995).\n14) A. Kitazawa and K. Nomura, J. Phys. Soc. Jpn. 66, 3944\n(1997).\n5" }, { "title": "2212.04423v1.Strong_photon_magnon_coupling_using_a_lithographically_defined_organic_ferrimagnet.pdf", "content": "Strong photon-magnon coupling using a lithographically de\fned\norganic ferrimagnet\nQin Xu,1Hil Fung Harry Cheung,1Donley S. Cormode,2Tharnier O. Puel,3Huma Yusuf,2\nMichael Chilcote,4Michael E. Flatt\u0013 e,3Ezekiel Johnston-Halperin,2and Gregory D. Fuchs4\n1Department of Physics, Cornell University, Ithaca NY 14853\n2Department of Physics, The Ohio State University, Columbus, OH 43210\n3Department of Physics and Astronomy,\nUniversity of Iowa, Iowa City, IA 52242\n4School of Applied and Engineering Physics,\nCornell University, Ithaca NY 14853\n(Dated: December 9, 2022)\nAbstract\nWe demonstrate a hybrid quantum system composed of superconducting resonator photons and\nmagnons hosted by the organic-based ferrimagnet vanadium tetracyanoethylene (V[TCNE] x). Our\nwork is motivated by the challenge of scalably integrating an arbitrarily-shaped, low-damping\nmagnetic system with planar superconducting circuits, thus enabling a host of quantum magnonic\ncircuit designs that were previously inaccessible. For example, by leveraging the inherent properties\nof magnons, one can enable nonreciprocal magnon-mediated quantum devices that use magnon\npropagation rather than electrical current. We take advantage of the properties of V[TCNE] x,\nwhich has ultra-low intrinsic damping, can be grown at low processing temperatures on arbitrary\nsubstrates, and can be patterned via electron beam lithography. We demonstrate the scalable,\nlithographically integrated fabrication of hybrid quantum magnonic devices consisting of a thin-\n\flm superconducting resonator coupled to a low-damping, thin-\flm V[TCNE] xmicrostructure.\nOur devices operate in the strong coupling regime, with a cooperativity as high as 1181(44) at\nT\u00180.4 K, suitable for scalable quantum circuit integration. This work paves the way for the\nexploration of high-cooperativity hybrid magnonic quantum devices in which magnonic circuits\ncan be designed and fabricated as easily as electrical wires.\n1arXiv:2212.04423v1 [quant-ph] 8 Dec 2022Hybrid quantum systems are attractive for emerging quantum technologies because they\ntake advantage of the distinct properties of the constituent excitations [1, 2]. This is impor-\ntant because no single quantum system is ideal for every task, e.g. scalable quantum infor-\nmation processing, quantum sensing, long-lived quantum memory, and long-range quantum\ncommunication all have di\u000berent requirements. Some hybrid systems that have been ex-\nplored are microwave photons hybridized with spins [3{15], optical photons hybridized with\natomic degrees of freedom [16{19], and superconducting qubits hybridized with phonons [20{\n23]. In creating hybrid systems, it is advantageous to operate in the strong-coupling, low-loss\nregime, where the relaxation rates of the two distinct quantum systems are exceeded by the\ncoupling rate between them. This allows the hybrid system to operate as a quantum in-\nterconnect, wherein quantum information can be passed from one excitation to another [2].\nThus, a central challenge is to couple distinct quantum systems strongly, with all elements\nmaintaining long coherence times. An equally critical challenge is to fabricate the hybrid\nquantum devices using scalable and integrable approaches so that their engineered properties\ncan be used in applications.\nCombining magnons (quantized spin waves) and microwave photons into a hybrid quan-\ntum system harnesses the unique properties of magnetic materials [24{27]. For example,\nmagnetic materials break time-reversal symmetry, providing a natural opportunity to create\nnonreciprocal devices. Circulators | classical but nonreciprocal circuit components based\non ferromagnetic resonance | are indispensable for superconducting qubits. Magnon-cavity\ninteractions provide new opportunities for nonreciprocal behavior [28, 29], which is useful\nalso for quantum devices [30, 31]. Additionally, magnetic materials are promising for cou-\npling to long-lived quantum spin systems [32{35], potentially enabling quantum intercon-\nnects between important quantum technologies based on microwave photons and on color\ncenter spins [36].\nThe challenge of integrating a microwave resonator with a low-damping magnetic ma-\nterial that can be lithographically patterned has limited the scalability of hybrid quantum\nmagnonic systems. Landmark initial demonstrations used large (millimeter-scale) crystals\nof the ferrimagnetic insulator yttrium iron garnet (YIG) because of its record-low damping,\neven at temperatures below 1 K for bulk crystals [37, 38]. These e\u000borts used either copper\nthree-dimensional microwave cavities [38{41] or a planar waveguide cavity [37]. With focused\nion beam milling and nanomanipulators, a micron-scale YIG crystal has been integrated with\n2a superconducting resonator [42]. Unfortunately, direct lithographic integration of thin-\flm\nYIG with superconducting planar circuits is an outstanding challenge, likely because lattice-\nmatched substrates such as gadolinium gallium garnet are strongly paramagnetic and thus\na lossy microwave substrate [43, 44]. Another approach was the integration of permalloy\n(Ni0:81Fe0:19), a ferromagnetic alloy that can be directly grown and patterned on a planar\ncavity without the need for high-temperature processing or lattice matching, which substan-\ntially enhances the scalability of this hybrid system [45, 46]. Unfortunately, permalloy has\norders-of-magnitude larger intrinsic magnetic damping as compared with high-quality bulk\nYIG crystals.\nIn this work, we demonstrate an alternate pathway to a strongly-coupled hybrid magnonic\nsystem in which a low-damping magnetic \flm is patterned directly on a superconducting\nresonator under gentle growth conditions. We develop a process to integrate lithographically-\npatterned \flms of the organic-based ferrimagnet vanadium tetracyanoethylene (V[TCNE] x)\nwith a planar superconducting microwave resonator. V[TCNE] xhas a typical Gilbert damp-\ning coe\u000ecient in the range \u000b= (4\u000020)\u000210\u00005[47{49], which is comparable to YIG bulk\ncrystals and high-quality YIG \flms grown on lattice matched substrates [50, 51]. We demon-\nstrate strong coupling between V[TCNE] xmagnons and resonator photons in two devices\n(3.6 GHz and 9.2 GHz) with a cooperativity as high as \u0018103. This is critically enabling\nfor integration and scaling, permitting future designs in which magnonic waveguides can be\ntailored as couplers or can mediate interactions between di\u000berent quantum excitations. Fo-\ncusing on the 3.6 GHz device, we present a detailed microwave transmission spectrum that\nreveals not only the expected avoided level crossing of the resonator mode and the uniform\nmagnon mode (the Kittel mode, or simply magnon mode unless otherwise stated) but also\nthe resonator mode hybridized with a discrete spectrum of excited magnon modes. Finally,\nwe study the relaxation rate of the hybrid system as a function of the system detuning\nin the frequency domain and the time domain. This work represents a paradigmatic shift\nin hybrid magnonic quantum systems by establishing an integrated and scalable platform\nenabling arbitrary design of the magnonic elements.\nThe Hamiltonian describing a magnon mode coupled to a resonator mode is [24, 38, 39,\n45, 52{54]:\nH0=~=!r\u0012\n^ay^a+1\n2\u0013\n+!m(B0)^by^b+g\u0010\n^by^a+^b^ay\u0011\n(1)\nwhere the \frst two terms describe, respectively, the energy of resonator photons and magnons\n3at a static \feld B0, while the third term describes the photon-magnon coupling. We have\nused ^aand ^ayto describe the creation and annihilation of resonator photons, and ^band^by\nto describe the creation and annihilation of magnons | these are derived from the Holstein-\nPrimako\u000b transformation on spin operators [55]. The collective coupling rate between the\nresonator photons and magnons is estimated as [38] g=gsp\nN, whereNis the number of\nspins within the magnetic material coupled to the resonator. One can estimate the single\nspin coupling rate from the geometry as [45, 56] gs=ge\u0016Bbrf!r=p8~Zr, wherebrf=\u00160=2w\nis the magnitude of the magnetic \feld experienced by V[TCNE] xspins per unit current in\nan inductor wire of width wwhen the spins are in close contact. We have also used the\nelectron Land\u0013 e gfactorge, the Bohr magneton \u0016B, and the characteristic impedance of the\nresonatorZr=p\nL=C.\nTo design a hybrid system that is useful for quantum circuit integration, we desire a sys-\ntem operating in the strong-coupling, low-loss regime in which both the resonator damping\nrate\u0014rand the magnon damping rate \u0014mare smaller than g. It is also useful to parameter-\nize the system in terms of its cooperativity, C= 4g2=\u0014r\u0014m, which exceeds 1 when the two\nsystems are strongly coupled [38]. In that case, when we tune the system into resonance,\nthe excitations are best described as hybrids of resonator photons and magnons.\nFirst we discuss the requirements of the magnetic material. Low intrinsic (Gilbert) damp-\ning is critical for minimizing \u0014m. Additionally, it is a major materials challenge to precisely\npattern and integrate low-damping magnetic material with the superconducting resonator.\nFor this purpose V[TCNE] xis advantageous because it can be grown on nearly any sub-\nstrate without the need for high temperature processing [57{60], giving it wide substrate\ncompatibility. Moreover, it can be patterned using electron beam lithography and lift-o\u000b\ntechniques without compromising the ultra-low damping [47, 61], which overcomes the chal-\nlenges of working with bulk-grown crystals and thus has considerable advantages for scaling\nand integration. One of the unusual properties of V[TCNE] xas a magnetic material is that\nit has a relatively low value of the saturation magnetization \u00160Ms\u001810 mT [47, 62, 63]. On\none hand, this could be a disadvantage in reaching a largep\nNto enable strong coupling.\nOn the other hand, it is advantageous from a device design point of view because it allows\none to work at comparatively small applied magnetic \feld, which avoids superconducting\nvortex formation.\nFor the microwave resonator we select a lumped-element design consisting of a planar in-\n4terdigitated capacitor that is shorted by a narrow inductor wire. This device is structurally\nsimilar to a transmon qubit that has a capacitor and a narrow inductor, except our structure\ndoes not include a Josephson junction as a part of the inductor. Instead, our narrow induc-\ntor wire e\u000eciently couples resonator excitations to the magnon mode when the magnetic\nmaterial is patterned directly on the wire surface. The resonator mode concentrates the\noscillating current through the wire, generating an Oersted magnetic \feld that excites the\nspins [14, 42, 45]. Using our device's characteristic impedance Zr=17.0(4.5) \n and inductor\nwidthw= 10\u0016m, we estimate gs= 36(5) Hz. Using N= 2:195\u00021012(from the magnetic\nvolume described below, and Ms), the total coupling rate is estimated to be g= 54(8) MHz.\nThe resonator is capacitively coupled to a coplanar microwave feedline that we use to excite\nand detect the coupled resonator-magnon system. Microwave electromagnetic simulations\nof the bare resonator design are available in the supplementary materials.\nThe fabrication procedure must integrate the growth and patterning of inorganic super-\nconducting \flms with organic-based V[TCNE] x[48, 64]. We begin by sputtering 60 nm of\nNb on a sapphire wafer and use photolithography and dry etching to form the microwave\nresonator. After wafer dicing, we spin a poly-methyl methacrylate (PMMA) bilayer on an\nindividual die, which we expose using ebeam lithography. Once developed and passivated,\nthe resulting structure is a lift-o\u000b template for CVD-deposited V[TCNE] x[47]. Inside an in-\nert gas glovebox, we deposit the V[TCNE] xat\u001860\u000eC [57, 58, 65] and then perform lift-o\u000b.\nThe V[TCNE] xis then encapsulated by epoxy and a glass coverslip, stabilizing the material\nfor weeks under ambient conditions [48, 64]. Figure 1(a) shows an optical micrograph of\nan encapsulated device with a resonator frequency !r=2\u0019\u00183.6 GHz. A zoom into the\nnarrow inductor shows the bright 10- \u0016m-wide Nb wire with a dark gray strip of patterned\nV[TCNE] xrunning down its center. The V[TCNE] xstrip is 300 nm thick, 6 \u0016m wide and\n600\u0016m long.\nWe start the electrical characterization of the device at 0.0809 T, a magnetic \feld at which\nthe resonator and magnons are strongly detuned. We measure the microwave transmission\nthrough the feedline using a vector network analyzer (VNA) at sample temperature T=\n0.43 K. The result is shown in Fig. 1(b). The transmission coe\u000ecient of a microwave\nresonator coupled to a transmission line can be modeled as [66]\njS21(!)j=\f\f\f\fa\u0002\u0012\n1\u0000(Ql=jQcj)ei\u001e\n1 + 2iQl(!=!res\u00001)\u0013\f\f\f\f; (2)\n5FIG. 1. (a) Microscope image of the 3.6 GHz resonator after V[TCNE] xwas deposited on its\ninductor wire. The V[TCNE] xcan be seen as a dark gray strip on the bright Nb wire at the\ncenter of the structure. The curved line is the border of the encapsulating epoxy. (b) Transmission\nspectrum (blue) and \ft (black dashed curve) at 0.0809 T when the device is at 0.43 K.\nwhereais the attenuation coe\u000ecient, !res\u0019!ris the mode resonance frequency and Qlis\nthe loaded quality-factor that is used to calculate the total damping rate via \u0014l=!res=Ql.\njQcjis the magnitude of the coupling quality-factor that parameterizes the loss of photons\nfrom the resonator to the waveguide and \u001eis the phase of Qc.\nFitting to Eqn. 2 reveals Ql= 4302,jQcj= 11200, and !r=2\u0019= 3.604 GHz. Since these\nmeasurements are far detuned from the magnon resonance, the measured damping rate is\napproximately the resonator damping \u0014r=!r=Ql= 2\u0019\u00020:8377 MHz. As discussed below,\nwe \fnd that !rand\u0014rhave weak but non-zero magnetic \feld dependence. We attribute\nthis behavior to vortices in the superconducting \flm that can increase \u0014r[67, 68]. We model\nthese contributions phenomenologically, assuming that they vary linearly with magnetic\n\feld. As discussed below, we estimate \u0014r=2\u0019to be 0:902(32) MHz at the resonance \feld\nBreswhere!m(Bres) =!r.\nNext we tune the external magnetic \feld closer to the avoided level crossing between the\nmicrowave resonator mode and the magnon mode to study the coupling between the two\nsystems. The coupled eigenfrequencies are [37]\n!\u0006=!r+ \u0001=2\u0006p\n\u00012+ 4g2=2; (3)\nwhere \u0001 = !m(B0)\u0000!ris the system detuning, and the magnon frequency is !m(B0) =\n\rp\nB0(B0+\u00160Me\u000b) for the static magnetic \feld B0applied parallel to the long axis of the\n6V[TCNE] xstrip. Here Me\u000b=Ms\u0000Hkdescribes the di\u000berence between Msand the uniaxial\nanisotropy \feld Hkof V[TCNE] x. The transmission spectrum through the feedline will be\nmodi\fed by the resonator and its interactions with the magnons [25]. The total transmission\ndue to line attenuation and magnon coupling is [37]\nS21(!;B 0) =S21;0(!)\u0012\n1 +(\u0014ext=2)e\u0000i\u001e\ni(!\u0000!r)\u0000\u0014r=2 +g2fi[!\u0000!m(B0)]\u0000\u0014m=2g\u00001\u0013\n;(4)\nwhere\u0014extdescribes the coupling rate between the feedline and the resonator, and S21;0(!)\nis the background transmission.\nFigure 2(a) shows the corresponding experimental data \u0001 jS21j= 20log 10(jS21j=jS21;0j),\nmeasured at T= 0.43 K as a function of magnetic \feld and frequency. It shows a strong\navoided level crossing | direct evidence of strong coupling between resonator photons and\nmagnons. These data are acquired with a VNA using \u000075 dBm of microwave power at the\nsample, which is far below the resonator's nonlinear power level, corresponding to about\n3:9\u0002106resonator photons for large \u0001 [69] (see supplementary materials for details). At\neach \feld, we extract the resonance frequencies !\u0006and resonance linewidths \u0014\u0006of the upper\nand lower branch by \ftting to Eqn. 2.\nTo resolve the features at small \u0001 and to determine gand\u0014maccurately, we measure\nS21with \fner magnetic \feld steps. The data are presented in Fig. 2(b). Qualitatively, the\nstrongest features are the two hybrid photon-magnon branches dispersing as !\u0006given by\nEqn. 3. Additionally, we observe fainter modes that are linearly dispersing with a slope of\n28 GHz/T. In contrast to the uniform magnon mode, we attribute those to magnon modes\nwith a \fnite wavevector k[70{72], which we discuss in further detail below.\nFigure 2(c) shows a line cut of \u0001 jS21jas a function of frequency acquired at B0=Bres\n(e.g. \u0001 = 0), with corresponding \fts to Eqn. 2 for each branch. The two resonances are the\n!\u0006as de\fned by Eqn. 3. We note that although !\u0000has a symmetric resonance lineshape, !+\nhas a Fano-like lineshape. Nevertheless, we can use the splitting between these resonances\nto \fndg= [!+(Bres)\u0000!\u0000(Bres)]=2 = 2\u0019\u0002[90:31(8)] MHz. The damping rate of the two\nbranches,\u0014\u0006are related to resonator and magnon damping rates via [25]\n\u0014\u0006= [\u0014r=2 +\u0014m=2\u0007Imp\n(\u0000!r+i\u0014r=2 +!m\u0000i\u0014m=2)2+ 4g2]: (5)\nFrom the linewidths in Fig. 2(c) we \fnd \u0014\u0000(Bres)=2\u0019= 16:37(29) MHz and \u0014+(Bres)=2\u0019=\n15:15(17) MHz. In the strong coupling regime ( g\u001d\u0014r;\u0014m),\u0014+=\u0014\u0000= (\u0014r+\u0014m)=2.\n7FIG. 2. Microwave measurement of V[TCNE] xcoupled to a 3.6 GHz resonator at 0.43 K. (a)\n\u0001jS21jplotted as a function of magnetic \feld and frequency. The white dashed lines mark the\nresonator and the magnon mode frequencies in the case where they are not coupled. (b) \u0001 jS21j\nacquired in \fner \feld steps near resonance \feld. We attribute the faint diagonal lines with slope\n28 GHz/T to be k6= 0 magnon modes. (c) \u0001 jS21jline cut atB0=Bres(e.g. \u0001 = 0) showing the\ntwo branch resonances at !+and!\u0000. The dashed lines are \fts as discussed in the text.\nTherefore we take the average of experimental \u0014+;\u0014\u0000to estimate ( \u0014r+\u0014m)=2. Together\nwith\u0014r(Bres)=2\u0019= 0.902(32) MHz, we estimate \u0014m=2\u0019= 30:62(34) MHz. We \fnd that this\ndevice operates with C= 1181(44) and ful\flls g>\u0014r;\u0014m.\nHaving probed the system in the frequency domain, we now turn to measurements of\nsystem relaxation in the time domain. For this, we apply pulsed microwave excitation to the\nfeedline and detect the ring-up/-down response in the time domain using a homodyne circuit.\nThe full circuit and measurement protocol is discussed in the supplementary materials. The\nmicrowave pulses are long enough to excite the system into driven steady-state oscillations.\nWhen the microwave drive turns o\u000b suddenly, the system will continue to oscillate, however,\nit will do so at its natural frequency. Its relaxation will include both intrinsic relaxation\n8FIG. 3. Ring-down experiment with homodyne circuit for the 3.6 GHz sample at 0.43 K, \u000065 dBm\nexcitation power. (a) Schematic plot of how the homodyne detected signal (red) changes with\nrespect to the microwave excitation power at the sample (green). When the microwave excitation\npower turns on (o\u000b), system will ring up (down) to steady-state. (b) Amplitude vs time plot of ring-\ndown at 0.101 T for on resonance excitation. The \ftted voltage ring-down time is 170.0(2.6) ns.\n(c) Amplitude vs time plot with 5 MHz detuning. (d) Total system damping calculated from\nring-down experiments (orange) and VNA experiments (blue).\ninto the environment and relaxation through radiation into the feedline that we detect. We\namplify and mix this signal with the reference tone, and digitize the result. Figure 3(a)\nshows the schematic plot of microwave power at the sample and the oscilloscope detected\nvoltage vs time.\nFigure 3(b) shows a log-linear plot of amplitude relaxation after driving the system\nwith\u000065 dBm of microwave power at the resonance frequency !+atT= 0.43 K and\nB0= 0:101 T. Under these conditions, the hybrid state has more resonator character than\nmagnon character, however, !+\u0000!r= 2\u0019\u000243.9 MHz\u0018g=2, indicating substantial magnon\n9participation. We observe exponential relaxation with a voltage ring-down time constant\nof 170.0(2.6) ns. Therefore, the energy relaxation time constant is \u001c= 85:0(1:3) ns, corre-\nsponding to a decay rate of \u0014+=2\u0019= 1:872(29) MHz. We can also detune the microwave\nexcitation with respect to the hybrid mode. In Fig. 3(c) we plot ring-down data acquired\nat the same \feld, except using a driving frequency that is detuned by 5 MHz from !+=2\u0019.\nIn this case, the homodyne signal beats with respect to the reference oscillator, giving rise\nto a decaying sinusoidal response. We recover a decay rate of 1.924(28) MHz on top of the\n5 MHz beat, which is consistent with the on-resonance driving result.\nTo explore the relaxation rates at di\u000berent detunings \u0001, we perform on-resonance ring-\ndown measurements as a function of B0. We plot the resulting damping rates along with\nthe damping rates extracted from VNA full-width at half-maximum (FWHM) linewidth\nmeasurements vs B0in Fig. 3(d). We see excellent agreement between the two methods,\nindicating that our VNA measurements are acquired in the linear response regime. Fig-\nure 3(d) underscores that the magnon damping rate is one order of magnitude larger than\nthe bare resonator damping rate. The damping rate of the upper (lower) branch is increasing\nwhen the \feld is tuned closer to Bresfrom below (above), since the hybrid state has a larger\nparticipation from the magnon mode as \u0001 approaches 0. The upper branch damping rate\nat 0.0809 T and the lower branch damping rate at 0.1281 T are assumed to be the resonator\ndamping rates with no magnon contribution since these \felds are far away from Bres. By\nlinear interpolation, we estimate \u0014r=2\u0019to be 0:902(32) MHz at Bres, which we used above\nto calculateC.\nWe now turn our attention to the additional magnon modes that are evident in Fig. 2(b),\ni.e., the lines indicated by arrows within the avoided level crossing region. For reference, we\nplot a dashed line showing the uniform magnon mode position if it were not coupled to the\nresonator. It separates the additional modes into higher frequency (red arrows) and lower\n(white arrows) frequency modes, which likely have distinct origins. Here, we treat all magnon\nmodes as spinwave modes characterized by a wavevector kand quantized by the boundary\nconditions at the surfaces of the V[TCNE] x. The case k= 0 refers to the uniform magnon\nmode, while the additional modes are k6= 0 magnon modes. The width and length of the\nV[TCNE] xstrip are long enough that quantization constrained in those directions cannot be\nresolved; only modes that are quantized in the thickness direction ( jkj\u0011kperpendicular to\nthe magnetic \feld) are resolvable. We assign the features at frequencies above the dashed\n10FIG. 4. (a) Energy spectrum of Hj ii=~!ij iicolored by the coe\u000ecient j ii=\u0010\nc(i)\nr;:::\u0011\nrelated\nto the resonator. This simulation used realistic values as follows: L= 300 nm,\r=2\u0019= 28 GHz/T,\n\u00160Me\u000b= 53:7 mT,\u00152\nex= 0:25\u000210\u000016m2,!r=2\u0019= 3:593 GHz,g=2\u0019= 90 MHz, and gn=g=(n+1)\nforn= 1;:::; 4. (b) The results labeled by `magnons' repeat the plot (a), while the red lines include\nonly the coupling between the uniform magnon mode (Kittel mode) and the resonator mode (both\nindependently identi\fed by the dashed lines).\nline to be thickness quantized k6= 0 magnon modes. We do not have a de\fnitive assignment\nfor the modes lying at frequencies below the dashed line. A more complete discussion can\nbe found in the supplemental material.\nTo better understand the transmission spectrum, we now discuss a model for multiple\nmagnon modes coupled to the resonator. We extend the Hamiltonian H0given in Eqn. 1 by\nadding a termHsthat describes the k6= 0 magnon modes as\nHs=~=X\nn=1!n^sy\nn^sn+ X\nn=1gn(^sy\nn^a+ ^sn^ay)!\n: (6)\nWe introduce creation and annihilation operators ^ snand ^sy\nnfor thek6= 0 magnons, and\ntheir direct coupling to the resonator has strength gn.\nWe consider magnon modes described by dipole-exchange interactions [73]; their fre-\nquency dispersion is given in Ref. [74],\n!n=\rp\n(B0+\u00160Me\u000b\u00152\nexk2\nn) (B0+\u00160Me\u000b+\u00160Me\u000b\u00152\nexk2\nn): (7)\nThe quantization index n= 1;2;:::is along the thickness direction, where the wavevectors\nare constrained by knL=n\u0019. Settingn= 0 recovers the frequency of the uniform magnon\n11mode (or Kittel mode), while the extra terms are due to the presence of exchange interac-\ntions with amplitude described by the exchange-length constant \u0015ex. The expression above\nassumesB0is oriented in the plane of V[TCNE]x.\nMagnon modes with k6= 0 can be directly excited by the magnetic \feld generated by the\ninductor only if their spatially-averaged amplitude does not vanish; therefore, their coupling\nto the resonator is highly dependent on the spin-pinning boundary conditions [47, 75]. For\ninstance, there is no direct coupling for totally unpinned boundaries, while for complete\npinning only even n-index modes couple. Here, we consider an intermediate situation where\nallgnare allowed to exist, which can happen for partial pinning [76]. To demonstrate the\nessential features of this interaction, we chose gn=g=(n+1) as a qualitative description of the\nfeatures observed, which is likely an overestimate. We solve the equation Hj ii=~!ij ii\nas a function of B0and plot the results in Fig. 4(a). The two branches !\u0006represent the\nhybridized modes due to the strong uniform magnon-resonator coupling. The k6= 0 modes\nhave weak \u0001jS21j(Fig. 2) because the j iihave small resonator amplitudes, however, they\nbecome stronger as their frequencies approach the !+branch (Fig. 4(a)). We truncate at\nn= 4 capturing only spin states that lie within the avoided-crossing gap, however, we note\nthat the spacing between the k6= 0 magnon modes is sensitive to the parameter \u0015exthrough\na linear dependence on \u00152\nex. The coupling of the k6= 0 magnons with the resonator increases\nthe total gap between the two branches !\u0006, as shown in Fig. 4(b).\nThese advances in our understanding of the quantum magnonic properties of V[TCNE] x\nhighlight its potential for applications in quantum information devices and the substan-\ntial potential for further optimization. In this study, the uniform mode magnon damp-\ning rate\u0014matT= 0:43 K corresponds to a FWHM ferromagnetic resonance (FMR)\nlinewidth of 1.09 mT, which is comparable to the room temperature 0.57 mT linewidth\nof the \\witness\" sample that was deposited at the same time as the V[TCNE] xused in\nthis device and measured at \u00189.8 GHz. This points to the opportunity for improvement\nthrough V[TCNE] xgrowth and patterning considering that room temperature thin \flm\nV[TCNE] xFMR linewidths as narrow as 0.094 mT [62], and ebeam patterned V[TCNE] x\nFMR linewidths as narrow as 0.12 mT, at 9.8 GHz have been demonstrated [47]. More-\nover, prior measurements of the FMR linewidth at temperatures down to 5 K revealed a\nstrong temperature-dependent strain e\u000bect, highlighting one avenue for improvement [63].\nV[TCNE] xis a promising material for quantum magnonic applications with properties that\n12rival and exceed that of YIG, is compatible with a broad range of materials, and can be\nlithographically integrated with planar superconducting circuits.\nIn conclusion, we have demonstrated a hybrid quantum system composed of supercon-\nducting resonator photons and magnons in the strong coupling regime with C= 1181. A\nkey advance of this work is the integration of lithographically patterned and low-damping\nmagnetic material with a superconducting circuit platform, enabling new quantum tech-\nnologies in which the electrical and the magnonic structures are designed and fabricated\non an equal basis. This capability can lead to nonreciprocal quantum circuit elements,\nnew forms of hybrid-system couplers, new opportunities for tunability, and the creation\nof transmon qubits with integrated magnonic properties. Beyond quantum technologies,\nthis hybrid system can o\u000ber sensitive readout of coherent magnonic excitations. The broad\ndesign space o\u000bered by lithographic patterning allows experiments that selectively probe dif-\nferent magnon wavevectors and { in combination with superconducting qubits { enables the\ncreation, manipulation, and detection of single magnons in a scalable, integrated platform.\nACKNOWLEDGEMENTS\nWe thank Brendan McCullian for useful conversations. The resonator design and fabri-\ncation, V[TCNE] xgrowth and growth optimization, and theory of uniform magnon mode\ncoupling were supported through the Center for Molecular Quantum Transduction (CMQT),\nan Energy Frontier Research Center supported by the Department of Energy O\u000ece of Sci-\nence, Basic Energy Sciences (DE-SC0021314). The development and design of resonator-\nV[TCNE] xintegration and lithography, measurement techniques, and theoretical analysis\nofk6= 0 magnon modes were supported by the Department of Energy O\u000ece of Science,\nBasic Energy Sciences Quantum Information Sciences program (DE-SC0019250). All mea-\nsurements were done using the CMQT low temperature facility at Cornell. This work also\nmade use of facilities at the Cornell NanoScale Facility, an NNCI member supported by\nthe NSF (NNCI-2025233) and the Cornell Center for Materials Research Shared Facilities\nwhich are supported through the NSF MRSEC program (DMR-1719875). We acknowledge\nthe support of the NanoSystems Laboratory User Facility which is supported by the Center\nfor Emergent Materials, an NSF MRSEC (DMR-2011876).\n13[1] Z.-L. Xiang, S. Ashhab, J. Q. You, and F. Nori, Hybrid quantum circuits: Superconducting\ncircuits interacting with other quantum systems, Reviews of Modern Physics 85, 623 (2013).\n[2] A. A. Clerk, K. W. Lehnert, P. Bertet, J. R. Petta, and Y. Nakamura, Hybrid quantum\nsystems with circuit quantum electrodynamics, Nature Physics 16, 257 (2020).\n[3] J. Verd\u0013 u, H. Zoubi, C. Koller, J. Majer, H. Ritsch, and J. Schmiedmayer, Strong magnetic\ncoupling of an ultracold gas to a superconducting waveguide cavity, Physical Review Letters\n103, 043603 (2009).\n[4] D. Schuster, A. Sears, E. Ginossar, L. DiCarlo, L. Frunzio, J. Morton, H. Wu, G. Briggs,\nB. Buckley, D. Awschalom, et al. , High-cooperativity coupling of electron-spin ensembles to\nsuperconducting cavities, Physical Review Letters 105, 140501 (2010).\n[5] Y. Kubo, F. Ong, P. Bertet, D. Vion, V. Jacques, D. Zheng, A. Dr\u0013 eau, J.-F. Roch, A. Au\u000b\u0012 eves,\nF. Jelezko, et al. , Strong coupling of a spin ensemble to a superconducting resonator, Physical\nReview Letters 105, 140502 (2010).\n[6] Y. Kubo, C. Grezes, A. Dewes, T. Umeda, J. Isoya, H. Sumiya, N. Morishita, H. Abe, S. On-\noda, T. Ohshima, et al. , Hybrid quantum circuit with a superconducting qubit coupled to a\nspin ensemble, Physical Review Letters 107, 220501 (2011).\n[7] R. Ams uss, C. Koller, T. N obauer, S. Putz, S. Rotter, K. Sandner, S. Schneider,\nM. Schramb ock, G. Steinhauser, H. Ritsch, et al. , Cavity QED with magnetically coupled\ncollective spin states, Physical Review Letters 107, 060502 (2011).\n[8] I. Chiorescu, N. Groll, S. Bertaina, T. Mori, and S. Miyashita, Magnetic strong coupling in a\nspin-photon system and transition to classical regime, Physical Review B 82, 024413 (2010).\n[9] P. Bushev, A. Feofanov, H. Rotzinger, I. Protopopov, J. Cole, C. Wilson, G. Fischer,\nA. Lukashenko, and A. Ustinov, Ultralow-power spectroscopy of a rare-earth spin ensemble\nusing a superconducting resonator, Physical Review B 84, 060501 (2011).\n[10] E. Abe, H. Wu, A. Ardavan, and J. J. Morton, Electron spin ensemble strongly coupled to a\nthree-dimensional microwave cavity, Applied Physics Letters 98, 251108 (2011).\n[11] X. Zhu, S. Saito, A. Kemp, K. Kakuyanagi, S.-i. Karimoto, H. Nakano, W. J. Munro,\nY. Tokura, M. S. Everitt, K. Nemoto, et al. , Coherent coupling of a superconducting \rux\nqubit to an electron spin ensemble in diamond, Nature 478, 221 (2011).\n14[12] J. Viennot, M. Dartiailh, A. Cottet, and T. Kontos, Coherent coupling of a single spin to\nmicrowave cavity photons, Science 349, 408 (2015).\n[13] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura,\nCoherent coupling between a ferromagnetic magnon and a superconducting qubit, Science\n349, 405 (2015).\n[14] A. Bienfait, J. J. Pla, Y. Kubo, M. Stern, X. Zhou, C. C. Lo, C. D. Weis, T. Schenkel, M. L. W.\nThewalt, D. Vion, D. Esteve, B. Julsgaard, K. M\u001clmer, J. J. L. Morton, and P. Bertet,\nReaching the quantum limit of sensitivity in electron spin resonance, Nature Nanotechnology\n11, 253 (2016).\n[15] D. Xu, X.-K. Gu, H.-K. Li, Y.-C. Weng, Y.-P. Wang, J. Li, H. Wang, S.-Y. Zhu, and\nJ. You, Quantum control of a single magnon in a macroscopic spin system, arXiv preprint\narXiv:2211.06644 (2022).\n[16] J. D. Thompson, T. Tiecke, N. P. de Leon, J. Feist, A. Akimov, M. Gullans, A. S. Zibrov,\nV. Vuleti\u0013 c, and M. D. Lukin, Coupling a single trapped atom to a nanoscale optical cavity,\nScience 340, 1202 (2013).\n[17] S. Spillane, T. Kippenberg, K. Vahala, K. Goh, E. Wilcut, and H. Kimble, Ultrahigh-Q toroidal\nmicroresonators for cavity quantum electrodynamics, Physical Review A 71, 013817 (2005).\n[18] A. Goban, C.-L. Hung, S.-P. Yu, J. Hood, J. Muniz, J. Lee, M. Martin, A. McClung, K. Choi,\nD. E. Chang, et al. , Atom{light interactions in photonic crystals, Nature Communications 5,\n3808 (2014).\n[19] H. J. Kimble, Strong interactions of single atoms and photons in cavity QED, Physica Scripta\n1998 , 127 (1998).\n[20] A. D. O'Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M. Lenander, E. Lucero, M. Nee-\nley, D. Sank, H. Wang, M. Weides, et al. , Quantum ground state and single-phonon control\nof a mechanical resonator, Nature 464, 697 (2010).\n[21] Y. Chu, P. Kharel, T. Yoon, L. Frunzio, P. T. Rakich, and R. J. Schoelkopf, Creation and\ncontrol of multi-phonon fock states in a bulk acoustic-wave resonator, Nature 563, 666 (2018).\n[22] K. J. Satzinger, Y. Zhong, H.-S. Chang, G. A. Peairs, A. Bienfait, M.-H. Chou, A. Cleland,\nC. R. Conner, \u0013E. Dumur, J. Grebel, et al. , Quantum control of surface acoustic-wave phonons,\nNature 563, 661 (2018).\n15[23] P. Arrangoiz-Arriola, E. A. Wollack, Z. Wang, M. Pechal, W. Jiang, T. P. McKenna, J. D.\nWitmer, R. Van Laer, and A. H. Safavi-Naeini, Resolving the energy levels of a nanomechanical\noscillator, Nature 571, 537 (2019).\n[24] O. O. Soykal and M. E. Flatt\u0013 e, Strong \feld interactions between a nanomagnet and a photonic\ncavity, Physical Review Letters 104, 077202 (2010).\n[25] M. Harder, B. Yao, Y. Gui, and C.-M. Hu, Coherent and dissipative cavity magnonics, Journal\nof Applied Physics 129, 201101 (2021).\n[26] H. Y. Yuan, Y. Cao, A. Kamra, R. A. Duine, and P. Yan, Quantum magnonics: When magnon\nspintronics meets quantum information science, Physics Reports 965, 1 (2022).\n[27] J. Shim, S.-J. Kim, S. K. Kim, and K.-J. Lee, Enhanced magnon-photon coupling at the\nangular momentum compensation point of ferrimagnets, Physical Review Letters 125, 027205\n(2020).\n[28] Y.-P. Wang, J. W. Rao, Y. Yang, P.-C. Xu, Y. S. Gui, B. M. Yao, J. Q. You, and C.-M. Hu,\nNonreciprocity and unidirectional invisibility in cavity magnonics, Physical Review Letters\n123, 127202 (2019).\n[29] X. Zhang, A. Galda, X. Han, D. Jin, and V. M. Vinokur, Broadband nonreciprocity enabled\nby strong coupling of magnons and microwave photons, Physical Review Applied 13, 044039\n(2020).\n[30] P. Lodahl, S. Mahmoodian, S. Stobbe, A. Rauschenbeutel, P. Schneeweiss, J. Volz, H. Pichler,\nand P. Zoller, Chiral quantum optics, Nature 541, 473 (2017).\n[31] N. R. Bernier, L. D. T\u0013 oth, A. Koottandavida, M. A. Ioannou, D. Malz, A. Nunnenkamp,\nA. K. Feofanov, and T. J. Kippenberg, Nonreciprocal recon\fgurable microwave optomechan-\nical circuit, Nature Communications 8, 604 (2017).\n[32] D. R. Candido, G. D. Fuchs, E. Johnston-Halperin, and M. E. Flatt\u0013 e, Predicted strong coupling\nof solid-state spins via a single magnon mode, Materials for Quantum Technology 1, 011001\n(2020).\n[33] A. B. Solanki, S. I. Bogdanov, M. M. Rahman, A. Rustagi, N. R. Dilley, T. Shen, W. Tong,\nP. Debashis, Z. Chen, J. Appenzeller, et al. , Electric \feld control of interaction between\nmagnons and quantum spin defects, Physical Review Research 4, L012025 (2022).\n[34] E. Lee-Wong, R. Xue, F. Ye, A. Kreisel, T. van Der Sar, A. Yacoby, and C. R. Du, Nanoscale\ndetection of magnon excitations with variable wavevectors through a quantum spin sensor,\n16Nano Letters 20, 3284 (2020).\n[35] B. A. McCullian, A. M. Thabt, B. A. Gray, A. L. Melendez, M. S. Wolf, V. L. Safonov,\nD. V. Pelekhov, V. P. Bhallamudi, M. R. Page, and P. C. Hammel, Broadband multi-magnon\nrelaxometry using a quantum spin sensor for high frequency ferromagnetic dynamics sensing,\nNature Communications 11, 5229 (2020).\n[36] D. Awschalom, K. K. Berggren, H. Bernien, S. Bhave, L. D. Carr, P. Davids, S. E. Economou,\nD. Englund, A. Faraon, M. Fejer, S. Guha, M. V. Gustafsson, E. Hu, L. Jiang, J. Kim,\nB. Korzh, P. Kumar, P. G. Kwiat, M. Lon\u0014 car, M. D. Lukin, D. A. B. Miller, C. Monroe,\nS. W. Nam, P. Narang, J. S. Orcutt, M. G. Raymer, A. H. Safavi-Naeini, M. Spiropulu,\nK. Srinivasan, S. Sun, J. Vu\u0014 ckovi\u0013 c, E. Waks, R. Walsworth, A. M. Weiner, and Z. Zhang,\nDevelopment of quantum interconnects (quics) for next-generation information technologies,\nPRX Quantum 2, 017002 (2021).\n[37] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Marx, R. Gross, and S. T. B.\nGoennenwein, High cooperativity in coupled microwave resonator ferrimagnetic insulator hy-\nbrids, Physical Review Letters 111, 127003 (2013).\n[38] Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, Hybridizing\nferromagnetic magnons and microwave photons in the quantum limit, Physical Review Letters\n113, 083603 (2014).\n[39] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Strongly coupled magnons and cavity mi-\ncrowave photons, Physical Review Letters 113, 156401 (2014).\n[40] J. A. Haigh, N. J. Lambert, A. C. Doherty, and A. J. Ferguson, Dispersive readout of ferro-\nmagnetic resonance for strongly coupled magnons and microwave photons, Physical Review\nB91, 104410 (2015).\n[41] Y. P. Wang, G. Q. Zhang, D. K. Zhang, X. Q. Luo, W. Xiong, S. P. Wang, T. F. Li, C. M.\nHu, and J. Q. You, Magnon kerr e\u000bect in a strongly coupled cavity-magnon system, Physical\nReview B 94, ARTN 224410 10.1103/PhysRevB.94.224410 (2016).\n[42] P. G. Baity, D. A. Bozhko, R. Mac^ edo, W. Smith, R. C. Holland, S. Danilin, V. Seferai,\nJ. Barbosa, R. R. Peroor, S. Goldman, U. Nasti, J. Paul, R. H. Had\feld, S. McVitie, and\nM. Weides, Strong magnon{photon coupling with chip-integrated yig in the zero-temperature\nlimit, Applied Physics Letters 119, 033502 (2021).\n17[43] M. Mary\u0014 sko, Paramagnetic resonance losses in GGG (Gd3Ga5O12) substrates, Czechoslovak\nJournal of Physics B 39, 116 (1989).\n[44] M. Mary\u0014 sko, In\ruence of GGG substrate on fmr and magnetostatic wave propagation, Journal\nof magnetism and magnetic materials 101, 159 (1991).\n[45] J. T. Hou and L. Liu, Strong coupling between microwave photons and nanomagnet magnons,\nPhysical Review Letters 123, 107702 (2019).\n[46] Y. Li, T. Polakovic, Y.-L. Wang, J. Xu, S. Lendinez, Z. Zhang, J. Ding, T. Khaire, H. Saglam,\nR. Divan, J. Pearson, W.-K. Kwok, Z. Xiao, V. Novosad, A. Ho\u000bmann, and W. Zhang, Strong\ncoupling between magnons and microwave photons in on-chip ferromagnet-superconductor\nthin-\flm devices, Physical Review Letters 123, 107701 (2019).\n[47] A. Franson, N. Zhu, S. Kurfman, M. Chilcote, D. R. Candido, K. S. Buchanan, M. E. Flatt\u0013 e,\nH. X. Tang, and E. Johnston-Halperin, Low-damping ferromagnetic resonance in electron-\nbeam patterned, high-Q vanadium tetracyanoethylene magnon cavities, APL Materials 7,\n121113 (2019).\n[48] H. F. H. Cheung, M. Chilcote, H. Yusuf, D. S. Cormode, Y. Shi, S. Kurfman, A. Franson,\nM. E. Flatt\u0013 e, E. Johnston-Halperin, and G. D. Fuchs, Raman spectroscopy and aging of the\nlow-loss ferrimagnet vanadium tetracyanoethylene, The Journal of Physical Chemistry C 125,\n20380 (2021).\n[49] J. S. Bola, H. Popli, R. M. Stolley, H. Liu, H. Malissa, O. Kwon, C. Boehme, J. S. Miller,\nand Z. V. Vardeny, Fabrication method, ferromagnetic resonance spectroscopy and spintron-\nics devices based on the organic-based ferrimagnet vanadium tetracyanoethylene, Advanced\nFunctional Materials 31, 2100687 (2021).\n[50] B. Lenk, H. Ulrichs, F. Garbs, and M. M unzenberg, The building blocks of magnonics, Physics\nReports 507, 107 (2011).\n[51] M. Onbasli, A. Kehlberger, D. H. Kim, G. Jakob, M. Kl aui, A. V. Chumak, B. Hillebrands,\nand C. A. Ross, Pulsed laser deposition of epitaxial yttrium iron garnet \flms with low Gilbert\ndamping and bulk-like magnetization, APL Materials 2, 106102 (2014).\n[52] E. T. Jaynes and F. W. Cummings, Comparison of quantum and semiclassical radiation the-\nories with application to the beam maser, Proceedings of the IEEE 51, 89 (1963).\n[53] M. Tavis and F. W. Cummings, Exact solution for an n-molecule|radiation-\feld hamiltonian,\nPhysical Review 170, 379 (1968).\n18[54] M. Tavis and F. W. Cummings, Approximate solutions for an n-molecule-radiation-\feld hamil-\ntonian, Physical Review 188, 692 (1969).\n[55] A. Aharoni, lntroduction to the Theory of Ferromagnetism (Oxford university press, 1996).\n[56] C. Eichler, A. Sigillito, S. A. Lyon, and J. R. Petta, Electron spin resonance at the level of\n10000 spins using low impedance superconducting resonators, Physical Review Letters 118,\n037701 (2017).\n[57] M. Harberts, Y. Lu, H. Yu, A. J. Epstein, and E. Johnston-Halperin, Chemical vapor deposi-\ntion of an organic magnet, vanadium tetracyanoethylene, Journal of Visualized Experiments\n, e52891 (2015).\n[58] N. Zhu, X. Zhang, I. Froning, M. E. Flatt\u0013 e, E. Johnston-Halperin, and H. X. Tang, Low loss\nspin wave resonances in organic-based ferrimagnet vanadium tetracyanoethylene thin \flms,\nApplied Physics Letters 109, 082402 (2016).\n[59] D. de Caro, M. Basso-Bert, J. Sakah, H. Casellas, J.-P. Legros, L. Valade, and P. Cassoux,\nCvd-grown thin \flms of molecule-based magnets, Chemistry of Materials 12, 587 (2000).\n[60] K. I. Pokhodnya, A. J. Epstein, and J. S. Miller, Thin-\flm V[TCNE] xmagnets, Advanced\nMaterials 12, 410 (2000).\n[61] N. Zhu, A. Franson, S. Kurfman, M. Chilcote, D. R. Candido, K. E. Nygren, M. E. Flatt\u0013 e,\nK. S. Buchanan, E. Johnston-Halperin, and H. X. Tang, Organic ferrimagnetic material vana-\ndium tetracyanoethylene for non-reciprocal microwave applications, in 2020 IEEE/MTT-S\nInternational Microwave Symposium (IMS) (IEEE, 2020) pp. 528{531.\n[62] A. H. Trout, S. W. Kurfman, Y. Shi, M. Chilcote, M. E. Flatt\u0013 e, E. Johnston-Halperin, and\nD. W. McComb, Probing the structure of vanadium tetracyanoethylene using electron energy-\nloss spectroscopy, APL Materials 10, 081102 (2022).\n[63] H. Yusuf, M. Chilcote, D. R. Candido, S. Kurfman, D. S. Cormode, Y. Lu, M. E. Flatt\u0013 e,\nand E. Johnston-Halperin, Exploring a quantum-information-relevant magnonic material: Ul-\ntralow damping at low temperature in the organic ferrimagnet V[TCNE] x, AVS Quantum\nScience 3, 026801 (2021).\n[64] I. H. Froning, M. Harberts, Y. Lu, H. Yu, A. J. Epstein, and E. Johnston-Halperin, Thin-\n\flm encapsulation of the air-sensitive organic-based ferrimagnet vanadium tetracyanoethylene,\nApplied Physics Letters 106, 10.1063/1.4916241 (2015).\n19[65] H. Yu, M. Harberts, R. Adur, Y. Lu, P. C. Hammel, E. Johnston-Halperin, and A. J. Epstein,\nUltra-narrow ferromagnetic resonance in organic-based thin \flms grown via low temperature\nchemical vapor deposition, Applied Physics Letters 105, 012407 (2014).\n[66] S. Probst, F. Song, P. A. Bushev, A. V. Ustinov, and M. Weides, E\u000ecient and robust analysis of\ncomplex scattering data under noise in microwave resonators, Review of Scienti\fc Instruments\n86, 024706 (2015).\n[67] D. F. Santavicca, J. K. Adams, L. E. Grant, A. N. McCaughan, and K. K. Berggren, Microwave\ndynamics of high aspect ratio superconducting nanowires studied using self-resonance, Journal\nof Applied Physics 119, 234302 (2016).\n[68] M. R. Vissers, J. Hubmayr, M. Sandberg, S. Chaudhuri, C. Bockstiegel, and J. Gao,\nFrequency-tunable superconducting resonators via nonlinear kinetic inductance, Applied\nPhysics Letters 107, 062601 (2015).\n[69] A. Schneider, Quantum Sensing Experiments with Superconducting Qubits (KIT Scienti\fc\nPublishing, 2021).\n[70] G. B. G. Stenning, G. J. Bowden, L. C. Maple, S. A. Gregory, A. Sposito, R. W. Eason, N. I.\nZheludev, and P. A. J. de Groot, Magnetic control of a meta-molecule, Optics Express 21,\n1456 (2013).\n[71] B. Bhoi, T. Cli\u000b, I. S. Maksymov, M. Kostylev, R. Aiyar, N. Venkataramani, S. Prasad, and\nR. L. Stamps, Study of photon{magnon coupling in a YIG-\flm split-ring resonant system,\nJournal of Applied Physics 116, 243906 (2014).\n[72] A. Serga, A. Chumak, and B. Hillebrands, YIG magnonics, Journal of Physics D: Applied\nPhysics 43, 264002 (2010).\n[73] A. G. Gurevich and G. A. Melkov, Magnetization Oscillations and Waves , 1st ed. (CRC Press,\n1996).\n[74] B. A. Kalinikos and A. N. Slavin, Theory of dipole-exchange spin wave spectrum for ferro-\nmagnetic \flms with mixed exchange boundary conditions, Journal of Physics C: Solid State\nPhysics 19, 7013 (1986).\n[75] H. Puszkarski, P. Tomczak, and H. T. Diep, Surface anisotropy energy in terms of magne-\ntocrystalline anisotropy \felds in ferromagnetic semiconductor (Ga,Mn)As thin \flms, Physical\nReview B 94, 195303 (2016).\n20[76] Q. Wang, B. Heinz, R. Verba, M. Kewenig, P. Pirro, M. Schneider, T. Meyer, B. L agel,\nC. Dubs, T. Br acher, and A. V. Chumak, Spin pinning and spin-wave dispersion in nanoscopic\nferromagnetic waveguides, Physical Review Letters 122, 247202 (2019).\n[77] P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau,\nE. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wilson,\nK. J. Millman, N. Mayorov, A. R. J. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, _I. Po-\nlat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen,\nE. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, P. van Mulbregt,\nand SciPy 1.0 Contributors, SciPy 1.0: Fundamental Algorithms for Scienti\fc Computing in\nPython, Nature Methods 17, 261 (2020).\n[78] W. Voesch, M. Thiemann, D. Bothner, M. Dressel, and M. Sche\u000fer, On-chip ESR measure-\nments of DPPH at mK temperatures, Physics Procedia 75, 503 (2015).\n21SUPPLEMENTARY MATERIALS\nI. DEVICE SIMULATIONS\nWe simulate the resonator with Keysight PathWave Advanced Design System (ADS)\nsoftware. Fig S1(a) shows the simulated transmission coe\u000ecient \u0001 jS21jvs frequency for the\n3.6 GHz resonator. The simulated resonance frequency is 3.933 GHz. We then simulate the\ndevice being driven at the resonance frequency and plot the time averaged magnitude of\ncurrent density in the superconducting \flm. The result is shown in \fg S1(b), in which red\nindicates larger current density and blue indicates smaller current density. This resonance\nmode has a large current density in the inductor wire for e\u000ecient coupling to magnon modes\nof a magnetic material deposited on the wire. After fabrication, we measure the resonator's\ntransmission spectrum using a VNA at 0 \feld before V[TCNE] xdeposition. The result is\nshown in \fg S1(c), where the resonance is at 3.804 GHz with Q-factor ( Ql) of 4922. This\nresonance frequency is lower than the simulation because of the \fnite kinetic inductance of\nNb [67, 68], which is not included in the simulation. This kinetic inductance adds to the\ngeometric inductance of the LC resonator and decreases the resonance frequency. Fig S1(d)\nshows the transmission spectrum at 0 \feld after V[TCNE] xdeposition and encapsulation.\nThe encapsulation epoxy and cover glass increase the e\u000bective dielectric constant of the\nresonator environment, which increases the capacitance C, resulting in a lower resonance\nfrequency. We speculate that the decrease of Q (2546) with epoxy encapsulation is caused by\nthe loss tangent of the epoxy and glass, and loss from the non-uniform V[TCNE] xmagnetiza-\ntion at 0 \feld. Then we increase the \feld to 0.0944 T to saturate V[TCNE] xmagnetization.\nFig 1(b) in the main text shows the resulting transmission spectrum. We attribute the\nincrease in Q (4302) to the increase of the uniformity of the V[TCNE] xmagnetization.\nII. DEVICE FABRICATION\nWe deposit Nb as the superconducting material for our device on a sapphire wafer by\nsputtering at 600 Celsius. The resulting Nb is 60 nm thick and has a critical temperature\n(Tc) of 7.3 K. The high T cof Nb and the low loss tangent of sapphire help limit the damping\nrate of the superconducting device.\nWe pattern the Nb \flm using photolithography followed by dry etching. We \frst spin-\n22FIG. S1. (a) Microwave transmission \u0001 jS21jfrom the ADS simulation of the 3.6 GHz resonator.\nWe \fnd a resonance frequency of 3.93 GHz. (b) The time averaged magnitude of current density\nplot of the resonator at the resonance frequency shown in (a). The zoom shows a large current\ndensity in the inductor wire. (c) Experimental measurement of \u0001 jS21jatB0= 0 for a resonator\nat 0.43 K without V[TCNE] x. Fitting with Eqn. 2 gives !r=2\u0019= 3:804 GHz and Ql= 4922. The\nresonance frequency is lower than the simulation because of the \fnite kinetic inductance of Nb in\nthe real device, which is not included in the simulation. (d) Experimental measurement of \u0001 jS21j\natB0= 0 for the V[TCNE] x-resonator device discussed in the main text, at 0.43 K. Fitting with\nEqn. 2 gives !r=2\u0019= 3:600 GHz and Ql= 2546.\ncoat the wafer with photoresist, expose using a 5x stepper, and then develop. Then we etch\nthe Nb using ion milling before stripping the photoresist.\nWe use an electron beam lithography lift-o\u000b process to pattern the V[TCNE] xas discussed\nin the main text. First we spin-coat individual resonator chips with a bilayer of PMMA\nresist and expose the resist layer. We then develop the PMMA layers, leaving an opening\n23for V[TCNE] xdeposition directly on the Nb. The V[TCNE] xis deposited using CVD as\ndiscussed in the main text. The PMMA is then dissolved using dichloromethane, which\nleaves the V[TCNE] xin the patterned regions untouched. The \fnal step is to encapsulate\nthe resonator and patterned V[TCNE] xusing epoxy and a coverslip. The \fnal dimensions\nof the V[TCNE] xis 600\u0016m\u00026\u0016m\u0002300 nm.\nIII. RESONATOR AND MAGNON MODE SATURATION POWER\nUsing power-dependent S21measurements at B0= 0:0994 T, we determine the saturation\npower (of the kinetic inductance non-linearity) for the 3.6 GHz resonator to be larger than\n\u000065 dBm applied to the feedline port. To make sure we are not saturating the magnetic\nresonance, we also make power-dependent S21measurements near Bres. No power depen-\ndence is observed in the range of \u000065 dBm to\u000085 dBm. The data in Fig. 2 are acquired\nwith\u000075 dBm of microwave power at the sample, which is well below saturation for both\nthe resonator and the magnet.\nWe now theoretically estimate the magnon precession cone angle. When we drive on\nresonance with the upper or lower branch !res, the average number of excitations in the\nresonator-magnet system is [69]\nhni=4Pin\n~!2\nresQ2\nl\njQcj;\nwherePin=\u000075 dBm = 3 :16\u000210\u000011W is the excitation power at the sample, Qlis the\nloaded Q-factor and Qcis the coupling Q-factor. For the upper branch at B0=Bres, we \fnd\nQl= 242:1,jQcj= 23000 and the resonance frequency !res=2\u0019= 3:669 GHz from \ftting\nto Eqn. 2. Under these conditions, we estimate hni= 5800 and the average number of\nmagnonshnmi=hni=2 = 2900 at B0=Bres.\nThe cone angle \u0012of the uniform magnon mode satis\fes\n1\u0000cos\u0012\u00191\n2\u00122=hnmi~!res\n1\n2N~!res;\nwhereN\u00192:195\u00021012is the estimated number of V[TCNE] xspins in the sample. Therefore,\nwe estimate \u0012\u00192p\nhnmi=N= 7:2\u000210\u00005rad = 0:0041\u000e.\nSimilarly, in Fig. 1(b) with !res=2\u0019= 3:604 GHz,Ql= 4302 andjQcj= 11200, we get\nthe average number of resonator photons hnri'hni= 3:9\u0002106.\n24IV. BACKGROUND TRANSMISSION\nThe background transmission S21;0(!) is assumed to be independent of B0and is measured\nat some (B0's,!'s) that are far away from the upper or lower branch (see Fig. 2(a), 2(b)).\nIn Fig. 2(a) for the 3.6 GHz device, S21;0(!) is measured from 3.48 GHz to 3.595 GHz at\n0.1024 T, and from 3.595 GHz to 3.71 GHz at 0.1064 T. In Fig. 2(b), S21;0(!) is measured\nfrom 3.395 GHz to 3.495 GHz and from 3.595 GHz to 3.795 GHz at 0.1071 T, and 3.495 GHz\nto 3.595 GHz at 0.1017 T.\nSimilarly, in Fig. 5(b) for the 9.2 GHz device, S21;0(!) is measured from 8.9 GHz to\n9.25 GHz at 0.2913 T, and from 9.25 GHz to 9.6 GHz at 0.3021 T. In Fig. 5(d), S21;0(!)\nis measured from 8.63 GHz to 9.23 GHz and from 9.47 GHz to 9.83 GHz at 0.2886 T, and\n9.23 GHz to 9.47 GHz at 0.3021 T.\nV. ERROR ANALYSIS\nUncertainty is calculated via standard error analysis from least-squared \ftting. We use\nthe python package \\scipy.optimize.curve \ft\" [77]. The output includes the optimized values\nof all the \ftting parameters and a 2-D array \\pcov\" which gives the covarience matrix. We\nreport the square root of the diagonal entries of the covariance matrix as the standard error.\nVI. EXTRACTED MAGNETIC PARAMETERS\nUsing the data shown in Fig. 2(a), we \ft !\u0006by \ftting to Eqn. 2. These results are then\n\ft to Eqn. 3 where !mis given by the Kittel formula with \r=2\u0019= 28 GHz/T. Also, we\nassume!r=!r0+\rrB0and treat!r0and\rrtogether with Me\u000bandgas free parameters.\nWe have used Me\u000b=Ms\u0000Hk, whereHkis the uniaxial anisotropy \feld. We obtain \u00160Me\u000b\n= 53.614(63) mT. Such a large Me\u000bis likely caused by the large strain applied to the\nV[TCNE] xdue to di\u000berential thermal expansion [63], which can induce a value of Hk>Ms.\nNext we determine the value of Bresusing the data shown in Fig. 2(b), again extracting\n!\u0006by \ftting to Eqn. 2. Figure S2 shows the extracted splitting !+\u0000!\u0000as a function of B0.\nFrom Eqn. 3 we know that !+\u0000!\u0000=p\n\u00012+ 4g2, where \u0001 = !m\u0000!r=\rrm\u0002(B0\u0000Bres),\nand\rrmis the change of !m\u0000!rper unit increase of B0. Treating g,Bresand\rrmas\nfree parameters, we get g=2\u0019= 90:43(8) MHz, Bres= 0:103429(18) T and \rrm=2\u0019=\n25FIG. S2. Upper and lower branch frequency di\u000berence vs \feld and the \ftting to extract gand\nBres. At the resonance \feld, the frequency di\u000berence is the smallest and is equal to 2 g.\n52:1(2:6) GHz/T. The best-\ft curve is shown as the black dashed line in Fig. S2. First, we\n\fnd a value of gthat is consistent with the value we extracted from the line cut shown in\nFig. 2(c). However, the \ftted \rrm=2\u0019is larger than the expected value of 28 GHz/T. Possible\nreasons for the discrepancy include (1) that interactions between the spin wave modes and\nthe upper (lower) branch distorts the shape of !+(B0>Bres) and!\u0000(B00 denotes the antiferromagnetic coupling between spins σi=±1/2 on\nthe sites of sublattice ’A’, and neighbouring spins Sj= 1,0,−1 on sites forming the\nsublattice ’B’. Then, spins on each sublattice tend to order ferroma gnetically, withMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 2\nopposite sign for the two types of spins. Dis the strength of a single–site anisotropy\n(or crystal–field) term acting only on the S=1 spins of sublattice B.\nOne may also choose σi=±1 rather than ±1/2. The change has to be taken\ninto account when calculating sublattice magnetizations and defining the compensation\npoint. Otherwise, themodifiedconvention simplyamountstoaresca lingoftheexchange\ncoupling [5, 10].\nThe mixed spin Ising model, eq. (1), is known, see, e.g., [3, 5, 10], to le ad to\ncompensation points for simple cubic lattices, in accordance with mea n–field theory [2],\nbut not for square lattices, in contrast to mean–field theory. As h as been observed\nalready some years ago by Buendia and Novotny [5, 12], compensat ion points may\noccur on square lattices, when adding a (ferromagnetic) coupling b etween next–nearest\nneighbouring (nnn) spins on the A sublattice. On the other hand, th e authors did not\nfind any evidence for compensation points, when considering nnn co uplings for B spins,\ninstead of the ones for A spins.\nIn the following article we shall challenge the latter suggestion, which seems to have\nbeen taken for granted by others, see, e.g., [7]. Indeed, based on extensive Monte Carlo\n(MC) simulations, we shall present clear evidence for compensation points due to nnn\nantiferromagnetic interactions between B, or S=1, spins on the sq uare lattice. In fact,\nthe effect is not easy to identify.\nThe outline ofthe article is as follows. Insection 2, to set the scene f or the following\nparts, we define the model with nnn couplings between S=1 spins, dis cuss its ground\nstates, and describe details of the simulations. The following section deals with the\ncompensation points. In section 4, MC results on critical propertie s and temperature\ndependent anomalies of the model will be presented. A brief summar y is given in the\nfinal section.\n2. Model, ground states, and simulations\nWe shall study the following mixed spin Ising model on a square lattice\nH=J1/summationdisplay\n/angbracketlefti,j/angbracketrightσiSj+D/summationdisplay\nj∈BSj2−J2/summationdisplay\n/angbracketlefti,j/angbracketright∈BSiSj (2)\nwith antiferromagnetic nearest neighbour couplings, J1>0 between A spins, σi=±1,\nand B spins, Sj= 0,±1, a crystal–field term of strength Dacting on the B spins, and\nnnn couplings, J2, between B spins. Note that we set A spins equal to ±1, in accordance\nwithprevious work [5, 10]. To identify possible compensation points, c are isthenneeded\nin defining the magnetization of the sublattice A, see above and below .\nTo determine the ground states of the model in the ( D/J1,J2/J1) plane, we first\nselect, like before [5, 12, 13], the structures which are stable amon g the ones described\nby 2×2 cells of spins on the square lattice. One readily observes the possib ility of\ndegenerate ground states, eventually with indefinitely large unit ce lls. Finally, we check\nthe tentative ground state structures by monitoring spin configu rations at very lowMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 3\ntemperatures obtained from careful cooling MC runs for lattices o f various sizes. The\nresulting ground state phase diagram is depicted in figure 1.\nThephasediagramatvanishingtemperature, T= 0,comprisesfourstructures. The\nantiferromagnetic(AF) structure isstableforferromagneticor weakly antiferromagnetic\nnnn couplings, J2, and negative or sufficiently small positive values of the crystal field ,\nD. The spins on each of the two sublattices order ferromagnetically, ±1, with opposite\nsign on the sublattices A and B.\nIn two of the remaining other three ground states, S=1 spins may b e in the state 0,\ndue to the single–site anisotropy term, D. Actually, for sufficiently large values of D, all\nspins of the sublattice B are in state 0. We abbreviate the resulting s tructure as ’0II’.\nThe structure is highly degenerate, with each A spin being either −1 or 1, leading to a\n2NA–fold degeneracy for NAsites on the sublattice A. In contrast, in the ’0I’ structure,\nsee figure 1, only half of the S=1 spins are in the state 0. They form, on the sublattice\nB, in the standard Wood’s notation for overlayers [14], a periodic c( 2×2) superlattice.\nThe other half of the spins on the B sites are ferromagnetically orde red,±1. The spins\non the sublattice A are ferromagnetically ordered as well, with oppos ite sign, compared\nto that of the B spins.\n-2-1012 3 4 56\nD/J1-3-2.5-2-1.5-1-0.500.51J2/J1AF\n0I 0II\nB-AF\nFigure 1. Ground state phase diagram of the mixed spin model, eq. (2), on a sq uare\nlattice, with solid lines separating the different phases. The dashed lin e in the 0I phase\nrefers to equation (7), see text.\nIn the fourth ground state structure, all spins on the sublattice B are\nantiferromagnetically aligned, due to the dominant antiferromagne tic nnn coupling, J 2.\nThe structure may beabbreviated asB–AF. Eachspin onthesublat tice Amay beeither\n−1 or 1, leading, as for 0II, to a large degeneracy.\nNote that there may be additional degeneracies at borderlines bet ween different\nground state structures [10].\nTo determine thermal properties of the model, (2), we mainly perfo rm MC\nsimulations, using the Metropolisalgorithmwithsingle–spin flips, provid ing, indeed, theMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 4\nrequired accuracy, so thatthereisno needtoapplyother techniq ues like cluster–updates\nor the Wang–Landau approach [15]. We consider lattices with LxLsites, employing full\nperiodic boundary conditions. Lranges from 4 to 120, to study finite–size effects.\nTypically, we do runs of 5 ×106to 107Monte Carlo steps per spin, where averages and\nerror bars may be obtained from evaluating a few of such runs, usin g different random\nnumbers. These quite long runs lead to reliable data, as before [10]. T he estimated\nerrors are usually smaller than the sizes of the symbols in the figures , and they are\nshown there only in a few cases.\nWe record the energy per site, E, the specific heat, C, both from the energy\nfluctuationsandfromdifferentiating Ewithrespecttothetemperature, andtheabsolute\nvalues of the sublattice magnetizations of the two sublattices\n|mA|=<|/summationdisplay\nAσi|> /(2(L2/2)) (3)\nand\n|mB|=<|/summationdisplay\nBSj|> /(L2/2) (4)\nas well as the absolute value of the staggered magnetization of sub lattice B, describing\nthe ordering for the B–AF structure,\n|mst\nB|=<|/summationdisplay\nB+Si−/summationdisplay\nB−Sj|> /(L2/2) (5)\nwith B+,−denoting an obvious bipartition of sublattice B. The brackets <>denote\nthe thermal average. Note the factor of 1/2 in the definition of |mA|, taking into\naccount the correct length of the S=1/2 spins, so that |mA(T= 0)|= 1/2 for the\nferromagnetic ground state of sublattice A. In addition, the corr esponding sublattice\n(staggered) susceptibilities, χA,χB, andχst\nB, have been computed from the fluctuations\nofthe(staggered)sublattice magnetizations. Wealsoanalyse the fourth–ordercumulant\nof various (sublattice) order parameters, the Binder cumulant [16 ], defined by\nU= 1−< m4> /(3< m2>2) (6)\nwith< m2>and< m4>being the second and fourth moment of (staggered)\nmagnetizationsofsublatticeAorB.Finally, wemonitortypicalequilibr iumMonteCarlo\nconfigurations, illustrating the microscopic behaviour of the syste m and providing, e.g.,\nevidence for the ground states structures, as mentioned above .\nTo test the accuracy of the simulations, we compared our MC data t o those of\nprevious accurate numerical work and to exact results by enumer ating all possible\nconfigurations for small lattices with L= 4 [5, 10].Mixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 5\n3. Compensation points\nCompensationpointsoccurattemperatures Tcomp,wherebothsublatticemagnetizations\ncancel each other, with a vanishing total magnetization. In finite s ystems, as one is\nstudying in MC simulations, a convenient and efficient way [5] to locate such points is to\nuse the crossing condition |mA|(Tcomp) =|mB|(Tcomp). At low temperatures sufficiently\nfar below the phase transition, finite–size effects are expected to be very weak, because\nthere the compensation effect is not related to critical phenomena .\n0 0.2 0.4 0.60.81 1.2 1.4 1.6\nkBT/J10.10.20.30.40.50.60.70.80.91|mA|, |mB|\nFigure 2. Sublattice magnetizations |mA|(open symbols, dotted lines) and |mB|(full\nsymbols,solidlines)atfixed D/J= 3.0andvarying J2/J1=0.25(circles), 0.0(triangles\ndown),−0.1 (squares), −0.24 (diamonds), and −0.26 (triangles left). Lattices of size\n602are simulated.\nIn the AF ground state, one has |mB|(T= 0)=1, while |mA|(T= 0)= 1/2. Now,\nat non–zero temperatures, |mB|(T) may fall off quite rapidly due to antiferromagnetic\ncouplings J2and due to a relatively large single–site anisotropy term, D, which favours\nflips of S=1 spins to state 0. Furthermore, at zero temperature, |mB|(T= 0) drops from\n1 to 1/2, when passing through the borderline between the AF and 0 I ground states.\nAccordingly, one might speculate that compensation points, at low t emperatures in the\nAF phase, may show up close to the AF–0I borderline at T=0. In fact, our preliminary\nmean–field calculations seem to suggest that lines of compensation p oints may spring\nfrom this AF–0I borderline.\nHowever, in our MC simulations, investigating carefully several case s, we find no\nevidence forsuch compensationpointsintheAFphase. Anexample is depicted infigure\n2. There, D/J1is fixed at 3.0, and J2/J1is varied, from +0.25 to −0.26, so that the\nAF–0I border at J2/J1=−0.25 is approached and crossed. One observes, that |mB|is\nalways larger than |mA|, albeit the difference between the two sublattice magnetizations\nmay get quite small when decreasing J2. MC data for lattices of fixed size, L= 60, areMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 6\ndisplayed. The observation on the absence of compensation points in the AF phase also\nholds for other lattice sizes as well as for other values of J2/J1andD/J1.\n0.10.2 0.3 0.4 0.50.60.7 0.8 0.9 1\nkBT/J10.10.20.30.40.50.60.70.80.91|mA|, |mB|\nTcomp\nFigure 3. Sublattice magnetizations |mA|(open symbols, dotted lines) and |mB|\n(full symbols, solid lines) at fixed J2/J1=−0.2, and varying D/J1= 3.0 (circles), 3.4\n(squares), 3.57 (diamonds), and 3.7 (triangles up), simulating lattic es of size L= 60.\nAtD/J1=3.57, a compensation point shows up, Tcomp.\nOn the other hand, compensation points will be argued below to occu r in the 0I\nphase, for J2/J1>−1.0, arising at vanishing temperature from the line\nJ2/J1=−2+D/(2J1) (7)\nwhich is depicted as the dashed line in the ground state phase diagram , figure 1.\nLet us first present numerical support for this claim, followed then by low\ntemperature energy considerations backing it up. A numerical exa mple is shown in\nfigure 3, where, for lattices with 602sites, the temperature dependences of the sublattice\nmagnetizations, |mA|and|mB|, are shown at fixed nnn coupling, J2/J1=−0.2, and\nvarying the strength of the single-site anisotropy term, D/J1, from 3.0 to 3.7. Note\nthat for this ratio of couplings, at T= 0, the AF–0I border is at 3.2, and the origin,\nat vanishing temperature, of the line of compensation points would b e, according to\nequation (7), at D/J= 3.6. As discussed in the context of figure 2, one observes\nthat|mB|(T) is always larger than |mA|(T) when being in the AF phase. However,\nin the 0I phase, here, at D/J1= 3.57, there seems to be a compensation point, with\nkBTcomp/J1≈0.26, clearly below the phase transition: At lower temperatures, |mB|is\nstill larger than |mA|, with reverse ordering at T > T comp. By further increase of D,\nD/J1= 3.7,|mA|(T) is larger than |mB|(T), at all temperatures up to Tc.\nTo establish numerically the presence of compensation points, care ful finite–size\nanalyses are required, as illustrated in figure 4. Here, MC data for v arious lattices sizes,\nwithLranging from 20 to 120, aredisplayed, at J2/J1=−0.2 andD/J1= 3.59, i.e. veryMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 7\n0.05 0.1 0.15 0.2 0.25 0.3\nkBT/J10.360.380.40.420.440.460.480.5|mA|, |mB|\nTcomp\nFigure 4. Sublattice magnetizations |mA|(open symbols, dotted lines) and |mB|\n(full symbols, solid lines) at J2/J1=−0.2 andD/J1= 3.59, with lattice sizes L= 20\n(squares), 40 (diamonds), 60 (triangles left), 80 (triangles down ), and 120 (circles).\nThe compensation point is located at kBT/J1≈0.16.\nclose to the point, from which, according to (7), the line of compens ation points may\narise. In fact, a compensation point may be located at kBTcomp/J1≈0.16. Below that\ntemperature, |mB|(T)supercedes |mA|(T), withfinite–size dependences being extremely\nsmall. Then, at Tcomp, a crossing of the two sublattice magnetizations occurs. The finite\nsize effects are still very weak up to kBT/J1≈0.25, providing clear evidence on the\nexistence of the compensation point in the thermodynamic limit.\nIndeed, similarfinite-sizeanalysesallowonetolocatetheratherste eplyrisinglineof\ncompensationpoints, atfixednnncouplings J2/J1=−0.2. Results onthecompensation\npoints,Tcomp, are summarized in figure 5, together with estimates for the trans ition line,\nTc, to the paramagnetic phase.\nThe critical line, Tc, has been determined by monitoring the size dependent\npositions of maxima in the specific heat C, in the susceptibility of the sublattice A,\nand in the intersection points of the Binder cumulant, especially, of s ublattice A. The\nintersections of the Binder cumulant U(L,T) for successive lattice sizes [16] turn out\nto have the smallest finite–size effects, thereby being most efficient in estimating the\ncritical temperature.\nThe line of compensation points seems to start at zero temperatur e atD/J1= 3.6\nforJ2/J1=−0.2, in agreement with equation (7). Furthermore, it extends only ov er\nquite a small region, with the compensation point coinciding with the cr itical point at\naboutD/J1= 3.52±0.02.\nLet us now turn to the low temperature considerations leading to (7 ). Specifically,\nwe consider, for the 0I structure, the one–spin flip excitations on sublattice B. For\nconcreteness, we analyse ground states with spins being in state −1 or 0 on sublatticeMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 8\n33.13.2 3.3 3.4 3.53.63.7 3.8 3.9 4\nD/J100.10.20.30.40.50.60.70.80.911.1kBT/J1\nAF 0ITc\nTcompparamagnetic\nFigure 5. Phase boundary to the paramagnetic phase, Tc(solid line), and\ncompensation points, Tcomp(dashed line), at J2/J1=−0.2, varying D/J1.\n22.12.2 2.3 2.4 2.52.62.7\nD/J100.10.20.30.40.50.60.70.80.9kBT/J1\n0Iparamagnetic\nTcompTc\nFigure 6. Phase boundary to the paramagnetic phase, Tc(solid line), and\ncompensation points, Tcomp(dashed line), at J2/J1=−0.8, varying D/J1.\nB, and in state 1 on sublattice A. Then there are four possible flips of S=1 spins,\nnamely flipping a spin from its state 0 to the state (i) −1 or (ii) 1, or flipping a spin\nfrom its state −1 to state (iii) 0 or (iv) +1. Obviously, only flip (i) will increase the\nsublattice magnetization |mB|above its value at zero temperature, |mB|(T= 0) = 1/2,\nwhile otherwise |mB|(T) will have a negative slope at low temperatures. Now, one\nmay readily calculate the energies of all four elementary flips. One ob tains that flip\n(i) costs the lowest energy either followed by flip (iii), when D/J1<2J2/J1+ 4,\nor followed by flip (ii), when J2/J1>−1.0. Equation (7) is then obtained underMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 9\nthe assumption, that an initial increase of |mB(T)|is needed to have a compensation\npoint, with |mB|>|mA|at sufficiently low temperatures. Otherwise, in the 0I phase,\n|mA|>|mB|at all temperatures 0 < T < T c, excluding a compensation point.\nWe further checked numerically the hypothesis, (7), in determining lines of\ncompensation points by fixing the nnn coupling J2/J1not only at −0.2, but also at\n−0.4 and−0.8, varying the strength of the single–site anisotropy term, D/J1. Indeed,\nwe observe compensation points in the 0I structure, which seem to arise, at zero\ntemperature, from D/J1= 3.2 at J2/J1=−0.4 and from D/J1= 2.4 at J2/J1=−0.8,\nin agreement with (7). The latter case is displayed in figure 6.\nNote that the range of values of D/J1, in which there is a line of compensation\npoints, shrinks appreciably as one decreases J2/J1. This may be seen by comparing\nfigures 5 and 6. Thence, it becomes increasingly difficult to locate com pensation points\nfor lower values of J2/J1.\n4. Phase transitions and anomalies\nWhile mixed spin Ising models are of much interest because of the pote ntial presence of\ncompensations points, they may exhibit other intriguing thermal pr operties as well.\nWe shall focus on two aspects, the characterization of the phase transitions\nassociatedwiththevariousgroundstates, andanomaliesinthetem peraturedependence,\nespecially, of the sublattice magnetization and the specifc heat C.\nOne expects that there is no phase transition associated with the h ighly degenerate\n0II ground state for the following reasons: At zero temperature , all spins of sublattice\nB are in state 0, while each spin of sublattice A may be either −1 or 1. The single–site\nanisotropy term, favouring the spin state 0 on sublattice B, does n ot support long–range\norder, in close analogy to the situation in the Blume–Capel model [17]. The spins on\nsublattice A act merely like a fluctuating random field, and may not lead to a phase\ntransition neither. In fact, our MC data, for various quantities, lik e the specific heat C\nor the probability to encounter a spin in state 0, give no indication for singular thermal\nbehaviour.\nIn case of the other three ground states, AF, 0I, and B–AF, we d etermine the\nuniversality class [18, 19] by analysing specific heat and susceptibilitie s. In particular,\nwemonitorthesize, L,dependenceofthatmaximumofthespecificheat, Cmax(L), which\ngoesover into a singularity inthe thermodynamic limit, L→ ∞(note that there may be\nother maxima as will be discussed below in the context of anomalies). F urthermore, we\nstudy the size dependence of the maxima in (sublattice) susceptibilit ies. For the AF and\n0I structures, we focus on χA,max(L). For the B–AF structure, we record the staggered\nsusceptibility of the antiferromagnetically ordered sublattice B, χst\nB,max(L). In all three\ncases, the critical behaviour is found to be consistent with having p hase transitions in\nthe universality class of the standard two–dimensional Ising model.\nTypical MC results on the size, L, dependence of maxima in the specific heat,\nCmax, are depicted in figure 7. For all three types of ground states, AF , 0I as wellMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 10\n2.2 2.4 2.62.8 3 3.2 3.4 3.63.8 4 4.2 4.4\nln(L)0.80.911.11.21.31.41.51.61.71.81.92Cmax\nFigure 7. Height of the critical maximum of the specific heat, Cmaxversus logarithm\nof the lattice size, ln L, forL= 10, 20, 40, 60, and 80 at (a) (circles) J2/J1=−0.3,\nD/J1=1.5,(b)(squares) J2/J1=−1.0,D/J1=1.5,and(c)(diamonds) J2/J1=−1.5,\nD/J1= 2.5. The three cases refer to the (a) AF, (b) 0I, and (c) B–AF st ructures.\nas B–AF, one approaches to a good degree, already for lattices of moderate sizes, the\nformCmax∝lnL, being characteristic for the two–dimensional Ising universality cla ss.\nNote that the prefactor in front of the logarithmic term depends s trongly on J2/J1and\nD/J1. Certainly, such Ising–like critical behaviour may be expected in the AF case. In\ncase of the B-AF phase, the spins on sublattice B order antiferrom agnetically, with the\nA spins providing effectively a randomly fluctuating field, which may be a rgued to be\nirrelevant for the universality class. Finally, in the 0I case, the Ising –like criticality may\nbe understood by the ferromagnetic order of the spins on sublatt ice A.\nThe analysis of the susceptibilies confirms the findings on the specific heat. Typical\nexamples, for the same model parameters as in figure 7, are shown in figure 8. The size\ndependent maxima in the susceptibilities are observed to follow the fo rmχmax∝L7/4,\nwith rather small corrections, for all three cases, AF, 0I, and B– AF. This behaviour\nmay be quantified by calculating the slope between successive points in the doubly\nlogarithmic plot of the MC data. The resulting effective local critical e xponent of the\nsusceptibility is near 7/4, even for moderate lattice sizes. Thence, in all three cases,\ncriticality seems to belong to the two–dimensional Ising universality c lass.\nCare is needed when attempting to determine the universality class f rom the value\nof the critical Binder cumulant U∗=U(Tc,L−→ ∞), because that value is known to\ndepend on boundary conditions, shape, and anisotropy of the cor relations [16, 20, 21].\nNotethatinthecases westudied, itsvalueseems tobeclose tothat ofthestandardtwo–\ndimensional Ising model with periodic boundary conditions for lattice s of square shape,\nU∗≈0.6107 [22]. However, the dependence, especially on anisotropy, may be very weak\n[20, 21], and a reliable analysis may require an exact or, at least, extr emely accurateMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 11\n2.2 2.4 2.62.8 3 3.2 3.4 3.63.8 4 4.2 4.4\nln(L)0123456ln(χA(B),max(st))\nFigure 8. Log–log plot for maxima of (staggered) sublattice susceptibilities χmax\nversus lattice size Lfor the AF, 0I, and B–AF cases, with the same parameters and\nnotation as in figure 7. In case of AF and 0I structures, χA,max, in case of the B–AF\nstructure, χst\nB,maxis recorded. For comparison, the solid line shows χmax∝L7/4.\ndetermination of the critical point. We therefore refrained from s uch an analysis here.\n0 0.2 0.4 0.60.81 1.2 1.4 1.6 1.8 2 2.2\nkBT/J100.10.20.30.40.5 C\nFigure 9. Specific heat Cas a function of temperature, kBT/J1, atJ2/J1=−0.2 and\nD/J= 3.4, simulating lattices of sizes L= 20 (circles), 40 (squares), 60 (diamonds),\nand 80 (triangles left).\nLet us now turn to discussing anomalies, exhibiting intriguing non–mon otonic\ntemperature dependences, which are not related to phase trans itions. Especially, we\nmonitor the magnetizations of sublattice B, |mB|, the specific heat C, and the thermallyMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 12\naveraged occupancy of B sites with spins in the state 0, pB(0).\nAn example for such anomalies is the overshooting of |mB|(T) at low temperatures\nin the vicinity of the 0I ground state, illustrated in figures 2 and 3. As discussed above,\nthe overshooting is related to decreasing the number of spins in sta te 0, by flipping such\nspins to, say, −1, costing minimal energy. Indeed, the anomaly in |mB|is monitored to\nbe accompanied by a pronounced decrease in pB(0).\nAnomalies may also occur in the specific heat, being caused either by a rapid\nincrease or decrease in the number of S=1 spins in state 0. For insta nce, such an\nanomaly in Chas been observed before [10] in the AF phase at J2/J1= 0, close to the\n0II structure. There, monitoring C(T), a three–peak–structure of the specific heat has\nbeen found, with a non-critical maximum at low temperatures, due t o easy flips of spins\nin the state 0, followed, at higher temperature, by a strongly size d ependent critical\npeak, and, at even larger temperature, another non–critical ma ximum, due to flipping\nof single spins for sublattices with rather large clusters of ’+’ or ’ −’ spins.\nWe observe a similar thermal behaviour of the specific heat in the 0I p hase, when\nturningonthennncoupling, J2<0. Anexample isdepicted infigure9, at J2/J1=−0.2\nandD/J1= 3.4. However, here the peak at lowest temperature, Tl, is due to a rather\ndrastic decrease of B spins in state 0, which has been discussed abo ve. Actually, by\nenhancing the strength of the single–site anisotropy term, D/J1= 3.6, the position of\nthat peak, at Tl, shifts first to somewhat higher temperatures. It still correspo nds to a\nlowering, with temperature, of the average number of 0’s for the s ublattice B, pB(0)(T).\nWhen moving even closer to the 0IIstructure, D/J1= 3.8 and 3.9, Tltends to be shifted\ntowards lower temperatures with increasing D. The peak position seems to follow the\nsame dependence as for vanishing J2[10], namely kBTl/J1∝4−D/J1. As may be seen\nfrom monitoring pB(0), the peak is then, as in the limit J2= 0, due to an increase in\nthe number of B spins in state 0 at low temperatures.\n5. Summary\nWehave studied a mixed spin Ising model with antiferromagneticcoup lings,J1, between\nspinsS=1/2andS=1onneighbouring sitesofasquarelattice, augme ntedbycouplings,\nJ2, between spins S= 1 on next–nearest neighbouring sites of the latt ice. An additional\nquadratic single–site anisotropy term, D, acts upon the S=1 spins. Based mainly on\ngroundstateconsiderationsandonextensive standardMonteCa rlosimulations, wehave\ndetermined the ground state phase diagram in the ( D/J1,J2/J1) plane and identified\ncompensation points, types of phase transitions corresponding t o different ground state\nstructures as well as anomalies for various physical quantities.\nIn particular, compensation points are found to exist, in contrast to previous belief.\nThey exist for antiferromagnetic nnn couplings, J2>−1.0, in the 0Iphase, springing, at\nzero temperature, from the line J2/J1=−2+D/(2J1). Across that line, different types\nof one–spin flips on the S=1 sublattice cost lowest energy, accompa nied by a change in\nthesignoftheslopeofthemagnetizationofthatsublatticeatlowte mperatures. AtfixedMixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 13\nvalue ofJ2/J1, the range, in D/J1, of compensation points is quite narrow, becoming\nextremely small for decreasing nnn couplings, as we have shown by d ecreasing J2/J1\nfrom−0.2 to−0.8.\nThe transition to the paramagnetic phase appears to be always in th e universality\nclass of the two–dimensional Ising model, with the critical maxima of t he specific\nheat growing with lattice size, L, in a logarithmic fashion, and with the maxima\nof the appropriate (staggered) sublattice susceptibilities growing with lattice size\nproportionally to L7/4. These results hold in case of AF, 0I, and B–AF ground states of\nthe model. In case of the 0II ground state, there is no phase tran sition.\nThe model is observed to exhibit interesting anomalies by showing non –monotonic\ntemperature dependences for various quantities, including sublat tice magnetization and\nspecific heat. The anomalies are typically caused by energetically fav oured flips of S=\n1 spins between the state 0 and a non–zero state.\nIn conclusion, our study provides clear evidence for non–expecte d compensation\npoints in a rather simple mixed Ising model with antiferromagnetic nex t–nearest\nneighbour couplings between S=1 spins on a square lattice. Extensio ns to three\ndimensional lattices are well beyond the present study, and may be investigated in\nthe future.\n6. Acknowledgements\nC E thanks the Department of Physics at the RWTH Aachen for the k ind hospitality\nduring his stay there. We thank Prof. Mark Novotny for a very use ful discusssion as\nwell as Dr. Mukul Laad for interesting conversations.\nReferences\n[1] N´ eel L 1948 Ann. Phys. (Paris) 3137\n[2] Kaneyoshi T and Chen J C 1991 J. Magn. Magn. Mater. 98201\n[3] Zhang G-M and Yang C-Z 1993 Phys. Rev. B489452\n[4] Bob´ ak A and Jascur M 1995 Phys. Rev. B5111533\n[5] Buendia G M and Novotny M 1997 J. Phys.: Condens. Matter 95951\n[6] Oitmaa J and Zheng W-H 2003 PhysicaA328185\n[7] Godoy M, Leite V S, and Figueiredo W 2004 Phys. Rev. B69054428\n[8] Jascur M and Strecka J 2005 Condens. Matter Phys. 8869\n[9] Oitmaa J and Enting I G 2006 J. Phys.: Condens. Matter 1810931\n[10] Selke W and Oitmaa J 2010 J. Phys.: Condens. Matter 22076004\n[11] Zukovic M and Bob´ ak A 2010 PhysicaA3895402\n[12] Buendia G M and Novotny M A 1997 J. Phys. IV (C1) France 7175\n[13] Buendia G M and Liendo J A 1997 J. Phys.: Condens. Matter 95439\n[14] Oura K, Lifshits V G, Saranin A A, Zotov A V and Katayama M 2003 Surface Science: An\nIntroduction (Berlin: Springer)\n[15] Landau D P and Binder K 2005 A Guide to Monte Carlo Simulations in Statistical Physics\n(Cambridge: Cambridge University Press)Mixed Ising ferrimagnets with next-nearest neighbour coup lings on square lattices 14\n[16] Binder K 1981 Z. Physik B43119\n[17] Blume M 1966 Phys. Rev. 141517; Capel H W 1966 Physica (Utr.) 32966\n[18] Fisher M E 1974 Rev. Mod. Phys. 46597\n[19] Pelissetto A and Vicari E 2002 Phys. Rep. 368549\n[20] Chen X S and Dohm V 2004 Phys. Rev. E70056136; Dohm V 2008 Phys. Rev. E77061128\n[21] Selke W and Shchur L N 2005 J. Phys. A: Math. Gen. 38L739; Selke W 2006 Eur. Phys. J. B\n51223\n[22] Kamieniarz G and Bl¨ ote H W J 1993 J. Phys. A: Math. Gen 26201" }, { "title": "1702.03609v4.Lieb_and_hole_doped_ferrimagnetism__spiral__resonating_valence_bond_states__and_phase_separation_in_large_U__AB__2___Hubbard_chains.pdf", "content": "Lieb and hole-doped ferrimagnetism, spiral, resonating valence-bond states, and phase separation\nin large-U AB 2Hubbard chains\nV . M. Martinez Alvarez1and M. D. Coutinho-Filho1\n1Departamento de F ´ısica, Laborat ´orio de F ´ısica Te ´orica e Computacional,\nUniversidade Federal de Pernambuco, Recife 50670-901, Pernambuco, Brazil\nThe ground state (GS) properties of the quasi-one-dimensional AB2Hubbard model are investigated taking\nthe effects of charge and spin quantum fluctuations on equal footing. In the strong-coupling regime, we derive\na low-energy Lagrangian suitable to describe the ferrimagnetic phase at half filling and the phases in the hole-\ndoped regime. At half filling, a perturbative spin-wave analysis allows us to find the GS energy, sublattice\nmagnetizations, and Lieb total spin per unit cell of the effective quantum Heisenberg model, in very good\nagreement with previous results. In the challenging hole doping regime away from half filling, we derive the\ncorresponding t-JHamiltonian. Under the assumption that charge and spin quantum correlations are decoupled,\nthe evolution of the second-order spin-wave modes in the doped regime unveils the occurrence of spatially\nmodulated spin structures and the emergence of phase separation in the presence of resonating-valence-bond\nstates. We also calculate the doping-dependent GS energy and total spin per unit cell, in which case it is shown\nthat the spiral ferrimagnetic order collapses at a critical hole concentration. Notably, our analytical results in\nthe doped regime are in very good agreement with density matrix renormalization group studies, where our\nassumption of spin-charge decoupling is numerically supported by the formation of charge-density waves in\nanti-phase with the modulation of the magnetic structure.\nI. INTRODUCTION\nMuch attention has been given to quantum phase transi-\ntions [1, 2], which are phenomena characterized by the change\nof the nature of the ground state (GS) driven by a non-thermal\nparameter: pressure, magnetic field, doping, Coulomb repul-\nsion, or competitive interactions. In this context, the study\nof quasi-one-dimensional (quasi-1D) compounds with ferri-\nmagnetic properties [3, 4] has attracted considerable theoreti-\ncal and experimental interest because of their unique physical\nproperties and very rich phase diagrams. In particular, the GS\nof quasi-1D quantum ferrimagnets with AB2orABB0unit\ncell topologies (diamond or trimer chains) described by the\nHeisenberg or Hubbard models [5] exhibit unsaturated spon-\ntaneous magnetization, ferromagnetic and antiferromagnetic\nspin-wave modes, effect of quantum fluctuations, and field-\ndependent magnetization plateaus, among several other fea-\ntures of interest.\nOf special interest is the topological origin of GS mag-\nnetic long-range order associated with the unit cell structure\nof the lattice [5–12]. These studies have been motivated\nand supported by exact solutions and rigorous results [13–\n18]; in particular, at half filling, the total spin per unit cell\nobeys Lieb-Mattis [13] (Heisenberg model) or Lieb’s theo-\nrem [15] (Hubbard model). On the other hand, it has been\nverified that the ferrimagnetic GS of spin- 1=2Heisenberg and\nHubbard=t-J AB 2chains, under the effect of frustration [19–\n23] or doping [7, 11, 24, 25], are strongly affected by quan-\ntum fluctuations that might cause its destruction and the occur-\nrence of new exotic phases: spiral incommensurate (IC) spin\nstructures, Nagaoka ( U!1 ) and resonating-valence-bond\n(RVB) states, phase separation (PS), and Luttinger-liquid be-\nhavior. These features can enhance the phenomenology in\ncomparison with a linear chain, which is dominated by the\nnontrivial Luttinger-liquid behavior that exhibits fractional\nexcitations [26, 27], emergent fractionalized particles [28],and fractional-exclusion statistic properties [29] in the spin-\nincoherent regime [30]. In addition, investigations of transport\nproperties in AB2chains, and related structures, have also un-\nveiled very interesting features [31].\nOn the experimental side, studies [32–34] of the magnetic\nproperties of homometallic phosphate compounds of the fam-\nilyA3Cu3(PO 4)4(A = Ca ,Sr,Pb) suggest that in these\nmaterials the line of trimers formed by spin- 1=2 Cu+2ions\nantiferromagnetically coupled do exhibit ferrimagnetism of\ntopological origin. Further, compounds Ca3M3(PO 4)4(M =\nNi;Co) with a wave-like layer structure built by zigzag M-\nchains exhibit antiferromagnetic ordering ( M = Ni ) or param-\nagnetic behavior ( M = Co ) [35]. On the other hand, bimetal-\nlic compounds, such as CuMn (S 2C2O2)2\u00017:5H2O[36], can\nbe modeled [36–38] by alternate spin- 1=2- spin- 5=2chains\nand support interesting field-induced quantum critical points\nand Luttinger-liquid phase [37]. In addition, frustrated dia-\nmond (AB2topology) chains can properly model the com-\npound azurite, Cu3(CO 3)2(OH) 2, in which case the occur-\nrence of the 1=3magnetization plateau is verified at high\nfields [39] in agreement with topological arguments [40] akin\nto those invoked in the quantum Hall effect. The spin-1/2\ntrimer chain compound Cu3(P2O6OH) 2, with antiferromag-\nnetic interactions only, also display the 1/3 magnetization\nplateau [41]. Interestingly, it has been established that in azu-\nrite the magnetization plateau is a dimer-monomer state [42],\ni.e., the chain is formed by pairs of S= 1=2monomers and\nS= 0dimers, with a small local polarization of the diamond\nspins [43], in agreement with density functional theory [44].\nThese dimer-monomer states have been found previously in\nthe context of modeling frustrated AB2chains [45–47], and\nconfirmed through a modeling using quantum rotors [48]. In\ncontrast to azurite, whose dimers appear perpendicular to the\nchain direction, in the spin-1/2 inequilateral diamond-chain\ncompounds [49] A3Cu3AlO 2(SO 4)4(A = K;Rb;Cs), the\nmagnetic exchange interactions force the dimers to lie along\nthe sides of the diamond cells and the monomers form a 1DarXiv:1702.03609v4 [cond-mat.str-el] 13 Dec 20182\nHeisenberg chain. In fact, the low-energy excitations of these\nnew compounds have been probed and a Tomonaga-Luttinger\nspin liquid behavior identified [50]. It is worth mentioning\nthat strongly frustrated AB2chains can exhibit ladder-chain\ndecoupling [20], in which case the ladder is formed via the\ncoupling between dimer spins in neighboring AB2unit cells.\nOn the other hand, besides the above-mentioned quasi-\n1D compounds and related magnetic properties, consider-\nable efforts have been devoted to the study of supercon-\nductivity and intriguing magnetic/charge ordered phases in\ndoped materials [51, 52], in particular the formation of spin-\ngapped states in compounds such as the family of doped\n(La;Sr;Ca)14Cu24O41. This compound is formed by one-\ndimensional CuO 2diamond chains, (Sr;Ca)layers, and two-\nlegCu2O3ladders [53]. These results certainly stimulate\nexperimental and theoretical investigations of quasi-1D com-\npounds in the hole-doped regime, which is the main focus of\nour work, as described in the following.\nIn this work, we shall employ an analytical approach suit-\nable to describe the strongly coupled Hubbard model on doped\nAB2chains, which were the object of recent numerical studies\nthrough density matrix renormalization group (DMRG) tech-\nniques [25]. Our functional integral approach, combined with\na perturbative expansion in the strong-coupling regime, was\noriginally proposed to study the doped Hubbard chain [54],\nand later adapted to describe various doped-induced phase\ntransitions in the U=1AB2Hubbard chain [55]. In ad-\ndition, this approach was used to describe the doped strongly\ncoupled Hubbard model on the honeycomb lattice [56], whose\nresults are very rewarding, particularly those for the GS en-\nergy and magnetization in the doped regime, which com-\npare very well with Grassmann tensor product numerical stud-\nies [57].\nThe paper is organized as follows: in Sec. II we review the\nfunctional integral representation of the Hubbard Hamiltonian\nin terms of Grassmann fields (charge degrees of freedom) and\nspinSU(2)gauge fields (spin degrees of freedom). In Sec. III\nwe diagonalize the Hamiltonian associated with the charge de-\ngree of freedom and obtain a perturbative low-energy theory\nsuitable to describe the ferrimagnetic phase at half filling and\nthe phases in the hole-doped regimes. In Sec. IV, we show\nthat the resultant Hamiltonian at half filling and large-U maps\nonto the spin- 1=2quantum Heisenberg model. In this regime,\na perturbative series expansions in powers of 1=Sof the spin-\nwave modes is presented, which allows us to calculate the\nGS energy, sublattice magnetizations, and Lieb GS total spin\nper unit cell in very good agreement with previous estimates.\nIn Sec. V, we derive the low-energy effective t-JHamilto-\nnian, which accounts for both charge and spin quantum fluc-\ntuations. We also present the evolution of the second-order\nspin-wave modes, GS energy and total spin per unit cell un-\nder hole doping, thus identifying the occurrence of spatially\nmodulated spin structures, with non-zero and zero GS total\nspin, and phase separation involving the later spin structure\nand RVB states at hole concentration 1=3. Remarkably, these\npredictions are in very good agreement with the DMRG data\nreported in Ref. [25]. Lastly, in Sec. VI, we present a sum-\nmary and concluding remarks concerning the reported results.II. FUNCTIONAL-INTEGRAL REPRESENTATION\nThe Hamiltonian of the one-band Hubbard model on chains\nwithAB2unit cell topology is given by [7, 8, 10]:\nH=\u0000X\nhi\u000b;j\fi\u001bft\u000b\f\nij^cy\ni\u000b\u001b^cj\f\u001b+H.c.g+UX\ni\u000b^ni\u000b\"^ni\u000b#;(1)\nwherei= 1;:::;Nc(=N=3)is the specific position of the\nunit cell, whose length is set to unity, Nc(N) is the num-\nber of cells (sites), \u000b;\f=A; B 1; B2denote the type of site\nwithin the unit cell, ^cy\ni\u000b\u001b(^ci\u000b\u001b) is the creation (annihilation)\noperator of electrons with spin \u001b(=\";#) at site\u000bof celli,\nand^ni\u000b\u001b= ^cy\ni\u000b\u001b^ci\u000b\u001bis the occupancy number operator. The\nfirst term in Eq. (1) describes electron hopping, with energy\nt\u000b\f\nij\u0011t, allowed only between nearest neighbors A-B1and\nA-B2linked sites of sublattices AandB(bipartite lattice),\nand the second one is the on-site Coulombian repulsive inter-\nactionU > 0, which contributes only in the case of double\noccupancy of the site i\u000b.\nAt this point, it is instructive to digress on some fundamen-\ntal aspects of the formalism used in our work [54–56]. With\nregard to the large-U doped Hubbard chain [54], U=1AB2\nHubbard chain [55] and the Hubbard model on the honeycomb\nlattice [56], it has been shown that the particle density product\nin Eq. (1) can be treated through the use of a decomposition\nprocedure, which consists in expressing ^ni\u000b\"^ni\u000b#in terms of\ncharge and spin operators:\n^ni\u000b\"^ni\u000b#=1\n2^\u001ai\u000b\u00002(^Si\u000b\u0001ni\u000b)2; (2)\nwhere\n^Si\u000b= 1=2X\n\u001b\u001b0^cy\ni\u000b\u001b0\u001b\u001b0\u001b^ci\u000b\u001b; (3)\nand\n^\u001ai\u000b= ^ni\u000b\"+ ^ni\u000b#; (4)\nare the spin-1/2 and charge-density operators, respectively,\n\u001b\u001b0\u001bdenotes the Pauli matrix elements (~\u00111), andni\u000bis\nan arbitrary unit vector. In fact, Eq. (2) follows from the iden-\ntity:1\n2^\u001ai\u000b\u0000^ni\u000b\"^ni\u000b#= 2( ^Sx;y;z\ni\u000b)2= 2( ^Si\u000b\u0001ni\u000b)2. The\nconvenience of using the decomposition defined in Eq. (2),\nwith explicit spin-rotational invariance for the large-U Hub-\nbard model, was discussed at length in Refs. [54–56].\nWe start by using the Trotter-Suzuki formula [58, 59],\nwhich allows us to write the partition function,\nZ=Tr[exp(\u0000\fH)], at a temperature kBT\u00111=\f, as\nZ=Trf^TQM\nr=1exp[\u0000\u000e\u001cH(\u001cr)]g, where ^Tdenotes the\ntime-ordering operator, the total imaginary time interval\nis formally sliced into Mdiscrete intervals of equal size\n\u000e\u001c=\u001cr\u0000\u001cr\u00001,r= 1;2;:::;M; with\u001c0= 0;and\n\u001cM=\f=M\u000e\u001c , under the limits M!1 and\u000e\u001c!0. We\nshall now introduce, between each discrete time interval, an\novercomplete basis of fermionic coherent states [58, 59], 1 =\u0001Q\ni\u000b\u001bdcy\ni\u000b\u001bdci\u000b\u001bexp\u0010\n\u0000P\ni\u000b\u001bcy\ni\u000b\u001bci\u000b\u001b\u0011\njfci\u000b\u001bgihfci\u000b\u001bgj,3\nwherefcy\ni\u000b\u001b;ci\u000b\u001bgdenotes a set of Grassmann fields satisfy-\ning anti-periodic boundary conditions: cy\ni\u000b\u001b(0) =\u0000cy\ni\u000b\u001b(\f)\nandci\u000b\u001b(0) =\u0000ci\u000b\u001b(\f); while the set of unit vectors\ndefines the vector field fni\u000bg, satisfying periodic ones:\nni\u000b(0) = ni\u000b(\f), under a weight functional (see below).\nThereby, following standard procedure [58, 59], the partition\nfunction reads:\nZ=\u0002Y\ni\u000b\u001bDcy\ni\u000b\u001bDci\u000b\u001bY\ni\u000bD2ni\u000bW(fni\u000bg)e\u0000\u0001\f\n0L(\u001c)d\u001c;\n(5)\nwhere the pertinent measures are defined by\nDcy\ni\u000b\u001bDci\u000b\u001b\u0011 lim\nM!1;\u000e\u001c!0M\u00001Y\nr=1dcy\ni\u000b\u001b(\u001cr)dci\u000b\u001b(\u001cr);(6)\nD2ni\u000b\u0011 lim\nM!1;\u000e\u001c!0M\u00001Y\nr=1d2ni\u000b(\u001cr); (7)\nthe weight functional, W(fni\u000bg), satisfies a normalization\ncondition at each discrete imaginary time \u001cr:\n\u0002Y\ni\u000bd2ni\u000bW(fni\u000b(\u001cr)g) = 1; (8)\nand the Lagrangian density L(\u001c)is written in the form:\nL(\u001c) =X\ni\u000b\u001bcy\ni\u000b\u001b@\u001cci\u000b\u001b\u0000X\nij\u000b\f\u001b(t\u000b\f\nijcy\ni\u000b\u001bcj\f\u001b+H.c.)\n+UX\ni\u000b[\u001ai\u000b\n2\u00002(Si\u000b\u0001ni\u000b)2]: (9)\nIn order to fix W(fni\u000bg)one should notice that, in the op-\nerator formalism: ^\u001a2\ni\u000b= ^\u001ai\u000b+ 2^ni\u000b\"^ni\u000b#. Therefore, using\nEq. (2), the following identity holds [54]:\n2(^Si\u000b\u0001ni\u000b)2=^\u001ai\u000b(2\u0000^\u001ai\u000b)\n2; (10)\nwhich means that the square of the spin component opera-\ntor along the ni\u000bdirection has zero eigenvalues if the site\nis vacant or doubly occupied, and a nonzero value only for\nsingly occupied sites, i.e., (^Si\u000b\u0001ni\u000b)2= 1=4. Now, taking\nadvantage of the choice of ni\u000b, the local spin-polarization and\nspin-quantization axes are both chosen along the ni\u000bdirec-\ntion. Therefore, for singly occupied sites, we find Si\u000b\u0001ni\u000b=\npi\u000b=2, withpi\u000b=\u00061, corresponding to the two possible\nspin-1/2 states. Further, by incorporating vacancy and double\noccupancy possibilities, corresponding to the four possible lo-\ncal states of the Hubbard model, one can write [54]\npi\u000b^Si\u000b\u0001ni\u000b=^\u001ai\u000b(2\u0000^\u001ai\u000b)\n2; (11)\nwithp2\ni\u000b= (\u00061)2. We stress that, due to fermion operator\nproperties, the square of Eq. (11) reproduces Eq. (10), and\na comparison between them implies, at arbitrary doping andU value, the formal equivalence between 2(^Si\u000b\u0001ni\u000b)2and\npi\u000b(^Si\u000b\u0001ni\u000b). In this context, we remark that the origi-\nnal Coulomb repulsion term of the Hubbard Hamiltonian in\nEq. (1) is formally and energetically (eigenvalues) equivalent\nto both that in Eq. (9) or in its linear version through the fol-\nlowing replacement: 2(^Si\u000b\u0001ni\u000b)2!pi\u000b(^Si\u000b\u0001ni\u000b). In-\ndeed, using the constraint in Eq. (11) we find, UP\ni\u000b[\u001ai\u000b\n2\u0000\npi\u000b(Si\u000b\u0001ni\u000b)] =UP\ni\u000b[\u001ai\u000b\n2\u00001\n2\u001ai\u000b(2\u0000\u001ai\u000b)], which is zero\nfor\u001ai\u000b= 0;1; whereas, as expected, for double occupied\nsites,\u001ai\u000b= 2, the local energy is U. Therefore, Eq. (11) in its\nGrassmann version, can be enforced by a proper choice of the\nnormalized weight functional:\nW(fni\u000bg) = lim\nM!1;\u000e\u001c!0MY\nr=1W(fni\u000b(\u001cr)g)\n=Cexpn\n\u0000\u0002\f\n0d\u001c\rX\ni\u000b[pi\u000bSi\u000b\u0001ni\u000b\u0000\u001ai\u000b\n2(2\u0000\u001ai\u000b)]2o\n;\n(12)\nwhere\r!1 in the continuum limit ( M!1 ,\u000e\u001c!0),\nwith delta-function peaks at the four local states of the Hub-\nbard model, andCis a normalization factor such that Eq. (8)\nholds. In fact, the product of W(fni\u000b(\u001cr)g)in Eq. (12) gener-\nates a sum in rin the exponential of the suitable chosen Gaus-\nsian function, i.e., W(fni\u000bg)is such that in the continuum\nlimit,M!1;\u000e\u001c!0, Eq. (12) obtains with a diverging \r,\nas pointed out in Ref. [54]. In this way, using Eq. (12) for the\nweight functional in Eq. (5) for the partition function Z, and\nintegrating overfni\u000bg, the Lagrangian density L(\u001c)in Eq.\n(9) can thus be written in the following linearized form [54]:\nL(\u001c) =X\ni\u000b\u001bcy\ni\u000b\u001b@\u001cci\u000b\u001b\u0000X\nij\u000b\f\u001b(t\u000b\f\nijcy\ni\u000b\u001bcj\f\u001b+H.c.)\n+UX\ni\u000b[\u001ai\u000b\n2\u0000pi\u000b(Si\u000b\u0001ni\u000b)]; (13)\nwhere the constraint in Eq. (11) was explicitly used.\nNow, since we are interested in studying the GS properties\nof theAB2Hubbard chains, we choose the staggered factor\npi\u000b= +1 (\u00001)at sites\u000b=B1; B2(A), consistent with the\nlong-range ferrimagnetic GS predicted by Lieb’s theorem at\nhalf filling and for any Uvalue [7, 8, 15], in which case we\nassume broken rotational symmetry along the z-axis. In this\ncontext, by considering the symmetry exhibited by the ferri-\nmagnetic order, let us define the SU(2)=U(1)unitary rotation\nmatrix [60]\nUi\u000b=2\n4cos\u0010\u0012i\u000b2\u0011\n\u0000sin\u0010\u0012i\u000b2\u0011\ne\u0000i\u001ei\u000b\nsin\u0010\u0012i\u000b2\u0011\nei\u001ei\u000b cos\u0010\u0012i\u000b2\u00113\n5;(14)\nwhere\u0012i\u000bis the polar angle between the z-axis and the unit lo-\ncal vector ni\u000band\u001ei\u000b2[0;2\u0019)is an arbitrary azimuth angle\ndue to theU(1)gauge freedom of choice for Ui\u000b. Moreover,\na new set of Grassmann fields, fay\ni\u000b\u001b;ai\u000b\u001bgcan be obtained,\naccording to the transformation:\nci\u000b\u001b=X\n\u001b0(Ui\u000b)\u001b\u001b0ai\u000b\u001b0; (15)4\nthat locally rotates each unit vector ni\u000bto thez-direction. On\nthe other hand, if we express the product \u001b\u0001ni\u000bin matrix\nform:\n\u001b\u0001ni\u000b=\u0014\ncos (\u0012i\u000b) sin (\u0012i\u000b)e\u0000i\u001ei\u000b\nsin (\u0012i\u000b)ei\u001ei\u000b\u0000cos (\u0012i\u000b)\u0015\n; (16)\nwe obtain, after using Eq. (14),\nUy\ni\u000b(\u001b\u0001ni\u000b)Ui\u000b=\u001bz; (17)\nwhich explicitly manifest the broken rotational symmetry\nalong thez-axis. In this way, by substituting Eqs. (14) and\n(15) into Eq. (3), and using the above result, we find\nSi\u000b\u0001ni\u000b=1\n2X\n\u001b\u001b0ay\ni\u000b\u001b[Uy\ni\u000b(\u001b\u0001ni\u000b)Ui\u000b]\u001b\u001b0ai\u000b\u001b0\n=1\n2X\n\u001b\u001b0ay\ni\u000b\u001b(\u001bz)\u001b\u001b0ai\u000b\u001b0\u0011Sz\ni\u000b; (18)\nthereby, the constraint in Eq. (11) can be written in the form\nSi\u000b\u0001ni\u000b=pi\u000b\u001ai\u000b(2\u0000\u001ai\u000b)\n2=1\n2(ay\ni\u000b\"ai\u000b\"\u0000ay\ni\u000b#ai\u000b#);\n(19)\nwherepi\u000b= +1 (\u00001)at sites\u000b=B1; B2(A). The choice\nofpi\u000babove implies Lieb’s ferrimagnetic ordering with the set\nf\u0012iA=\u0012iB1=\u0012iB2= 0g, for alli, at half filling. However,\nin the hole doped regime away from half filling, the \u0012i\u000b’s can\nbe nonzero (e.g., \u0012i\u000b=\u0019for a spin flip, leading to a change\nin the sign of Sz\ni\u000b); further,Sz\ni\u000bcan be zero either by the pres-\nence of holes or doubly occupied sites ( ay\ni\u000b\"ai\u000b\"=ay\ni\u000b#ai\u000b#).\nLastly, using Eqs. (15) and (19) into the Lagrangian, Eq. (13),\nwe find, after suitable rearrangement of terms,\nL(\u001c) =L0(\u001c) +Ln(\u001c); (20)\nwhere both Lagrangians are quadratic in the Grassmann fields:\nL0(\u001c) =X\ni\u000b\u001bay\ni\u000b\u001b@\u001cai\u000b\u001b\u0000X\ni\u000bj\f\u001b(t\u000b\f\nijay\ni\u000b\u001baj\f\u001b+H.c.)\n+U\n2X\ni\u000b\u001b(1\u0000pi\u000b\u001b)ay\ni\u000b\u001bai\u000b\u001b; (21)\nand\nLn(\u001c) =X\ni\u000b\u001b\u001b0ay\ni\u000b\u001b0(Uy\ni\u000b@\u001cUi\u000b)\u001b0\u001bai\u000b\u001b\n\u0000X\ni\u000bj\f\u001b\u001b0t\u000b\f\nij[ay\ni\u000b\u001b0(Uy\ni\u000bUj\f\u00001)\u001b0\u001baj\f\u001b+H.c.];(22)\nwith the first term in both Eqs. (21) and (22) being originated\nfrom the first term in Eq. (13), the second ones come from\nthe hopping term in Eq. (13), after a rearrangement of terms,\nwhile the last one in Eq. (21) (proportional to U) is obtained\nby using Eq. (19) in the last term of Eq. (13). It is worth\nmentioning that only charge degrees of freedom (Grassmann\nfields) appear inL0(\u001c), and spin degrees of freedom under theconstraint in Eq. (19) [ SU(2)gauge fieldsfUy\ni\u000b;Ui\u000bg, which\ncarry all the information on the vector field fni\u000bg] are now\nrestricted toLn(\u001c), which includes both spin and charge de-\ngrees of freedom.\nIn the large-U regime, double occupancy is energetically\nunfavorable and the factor 2\u0000\u001ai\u000bis no longer needed in\nEq. (19), i.e., Si\u000b\u0001ni\u000b=pi\u000b\u001ai\u000b\n2, with\u001ai\u000b= 0 or1. In\nthis case, a proper perturbative analysis will allow us to study\nhole doping effects in Sec. V in a macroscopic fashion, so we\ndefine\n\u000e= 1\u00001\nNX\ni\u000bh\u001ai\u000bi; (23)\nwhich measures the thermodynamic average of hole doping\naway from half filling. In this context (strong-coupling limit),\nwe take advantage of results derived from L0(\u001c)(charge ef-\nfects in Sec. III), and at half filling (Sec. IV), in which case\ncharge degrees of freedom are frozen.\nIII. CHARGE DEGREES OF FREEDOM AND THE\nSTRONG-COUPLING LIMIT\nIn this section, we shall first diagonalize the Hamiltonian\nassociated with the Lagrangian L0(\u001c)through the use of a\nspecial symmetry property of the AB2chains and a canonical\ntransformation in reciprocal space. Then, by introducing a\nperturbative expansion in the strong-coupling regime, a low-\nenergy effective Lagrangian for the AB2Hubbard chains at\nhalf filling and in the doped regime will be obtained.\nA. Charge degrees of freedom\nWe begin our discussion by considering the Lagrangian L0\nin Eq. (21), and its corresponding Hamiltonian H0, free of\ntheSU(2)gauge fields. By performing the Legendre trans-\nformation:H0=\u0000P\ni\u000b\u001b@L0\n@(@\u001cai\u000b\u001b)@\u001cai\u000b\u001b+L0, where\n@L0\n@(@\u001cai\u000b\u001b)=ay\ni\u000b\u001b, the resultingH0is given by\nH0=\u0000X\nhi\u000b;j\fi\u001b(t\u000b\f\nijay\ni\u000b\u001baj\f\u001b+H.c.)\n+U\n2X\ni\u000b\u001b(1\u0000pi\u000b\u001b)ay\ni\u000b\u001bai\u000b\u001b: (24)\nFurther, sinceH0(L0) is quadratic in the Grassmann fields,\nthe solution for the energy of the system is given by H0in its\ndiagonalized form [59].\nTheAB2unit cell topology exhibits a symmetry [9, 11, 24,\n25, 55] under the exchange of the labels of the Bsites in a\ngiven unit cell. Thus, we can construct a new set of Grass-\nmann fields possessing this symmetry, i.e., either symmet-\nric or antisymmetric with respect to the exchange operation\nB1$B2:\n(di\u001b;ei\u001b) =1p\n2(aiB1\u001b\u0006aiB2\u001b); bi\u001b=aiA\u001b: (25)5\nIn addition, as a signature of the quasi-1D structure of the\nAB2chains, we notice that the B1andB2sites are located at\na distance 1=2(in units of length) ahead of the Asite. There-\nfore, after Fourier transforming the above Grassmann fields,\ni.e.,fdi;\u001b;ei;\u001b;bi;\u001bg=1pNcP\nkeikxifdk;\u001b;ek;\u001b;bk;\u001bg, it is\nconvenient to introduce a phase factor eik\n2through the follow-\ning transformation [55]: (Ak\u001b;Bk\u001b) =1p\n2(dk\u001b\u0006eik\n2bk\u001b), so\nthatH0in Eq. (24) thus becomes\nH0=X\nk\u001b\"k[Ay\nk\u001bAk\u001b\u0000By\nk\u001bBk\u001b] +U\n2X\nk\u001b(1\u0000\u001b)ey\nk\u001bek\u001b\n+U\n2X\nk\u001b[Ay\nk\u001bAk\u001b+By\nk\u001bBk\u001b\u0000\u001b(Ay\nk\u001bBk\u001b+By\nk\u001bAk\u001b)];\n(26)\nwhere\n\"k=\u00002p\n2tcos(k=2); (27)\nwithk= 2\u0019j(3\nN)\u0000\u0019, andj= 1;:::;N= 3. We can now ex-\nactly diagonalizeH0through the following Bogoliubov trans-\nformation:\nAk\u001b=uk\u000bk\u001b\u0000\u001bvk\fk\u001b; Bk\u001b=\u001bvk\u000bk\u001b+uk\fk\u001b;(28)\nwithukandvksatisfying the canonical constraint: (uk)2+\n(vk)2= 1 , to maintain the anticommutation relations of\nthe Grassmann fields. Due to the ferrimagnetic order of the\nGS, the above transformation is subject to a 4\u0019periodicity\nof the Bogoliubov functions fuk;vkgand Grassmann fields\nf\u000bk\u001b;\fk\u001bg. The diagonalized H0thus reads:\nH0=\u0000X\nk\u001b(Ek\u0000U\n2)\u000by\nk\u001b\u000bk\u001b+X\nk\u001b(Ek+U\n2)\fy\nk\u001b\fk\u001b\n+U\n2X\nk\u001b(1\u0000\u001b)ey\nk\u001bek\u001b; (29)\nwhere\n(uk;vk) =1p\n2\u0012\n1\u0006j\"kj\nEk\u00131=2\n; (30)\nand\nEk=q\n\"2\nk+U2=4: (31)\nAs one can see from Eq. (29), the non-interacting tight bind-\ning (U= 0) spectrum ofH0present three electronic bands:\na nondispersive flat band (related to the Grassmann fields\nfey\nk\u001b;ek\u001bg, macroscopically degenerate), and two dispersive\nones. InAB2chains, flat bands are closely associated with\nferrimagnetism (unsaturated ferromagnetism) [5, 7, 8] at half\nfilling, in agreement with Lieb’s theorem [15, 16], or fully\npolarized ferromagnetism [17] associated with the flat lowest\nband. We also stress that even at this level of approximation\nand in the weak coupling regime ( U= 2t), it was shown [7]\nthat hole doping [parametrized by \u000edefined in Eq. (23)] candestroy the ferrimagnetic order and/or induce phase separa-\ntion inAB2chains. As depicted in Fig. 1(a), the U= 0 spin\ndegeneracy of the flat bands is removed by the Coulombian\nrepulsive interaction, in which case a gap Uopens between\ntheek\u001bmodes:ek\"= 0, where spins at sites B1andB2are\nup, andek#=U, where these spins are down. On the other\nhand, the two dispersive bands are spin degenerated, and also\ndisplay a Hubbard gap Useparating the low (\u000bk\u001b)-energy and\nhigh(\fk\u001b)-energy modes [55].\nB. Strong-coupling limit\nIn this subsection, we shall introduce a perturbative expan-\nsion in the strong-coupling regime ( U\u001dt) in order to obtain\na low-energy effective Lagrangian for the AB2Hubbard chain\nat half filling and in the doped regime. First, we resume the re-\nsults of the previous section by writing the Grassmann fields\ndi\u001bandbi\u001bin terms of the Grassmann (Bogoliubov) fields\n\u000bk\u001band\fk\u001b:\n(di\u001b;bi\u001b) =1p2NcX\nk(eikxi;eik(xi\u00001\n2))\n\u0002[(uk\u0006\u001bvk)\u000bk\u001b\u0006(uk\u0007\u001bvk)\fk\u001b];(32)\nwhere the phase factor e\u0000ik\n2signalizes the quasi-1D AB2\nstructure, and the antisymmetric Grassmann field ei;\u001bremains\nas defined in Eq. (25). In the strong-coupling limit, however,\nit will prove useful to define a set of auxiliary spinless Grass-\nmann fields [54, 55] in direct space associated with di\u001band\nbi\u001b:\n(\u000bi;\fi) =r\n1\nNcX\nk;\u001b\u0012(\u0006\u001b)eikxi(\u000bk\u001b;\fk\u001b); (33)\nand a similar equation for (\u000b1\n2\ni;\f1\n2\ni)$(\u000bk\u001b;\fk\u001b)is obtained\nby the replacements: \u0012(\u0006\u001b)!\u0012(\u0007\u001b)andxi!xi\u00001=2,\nwhere\u0012(\u001b)is the Heaviside function, while for the antisym-\nmetric component, one has\nei;\u001b=r\n1\nNcX\nkeikxiek;\u001b: (34)\nNow, by expanding (uk;vk)in Eq. (30) in powers of t=U:\n(uk;vk)\u00191p\n2\u0014\n1\u0006j\"kj\nU+O\u0012t2\nU2\u0013\u0015\n; (35)\nsubstituting these results into the Eq. (32), and using the in-\nverse transformation of Eq. (33), we can derive a perturbative\nexpansion in powers of t=U for the Grassmann fields di\u001band\nbi\u001bin terms of the spinless Grassmann fields as follows:\ndi\u001b=\u0012(\u001b)\u000bi+\u0012(\u0000\u001b)\fi+p\n2t\nU\u0012(\u0000\u001b)(\u000b1\n2\ni+\u000b1\n2\ni+1)\n+t\nU\u0012(\u001b)[p\n2(\f1\n2\ni+\f1\n2\ni+1)\u0000t\nU(2\u000bi+\u000bi+1+\u000bi\u00001)]\n+O(t2=U2); (36)6\nbi\u001b=\u0012(\u0000\u001b)\u000b1\n2\ni\u0000\u0012(\u001b)\f1\n2\ni+p\n2t\nU\u0012(\u001b)(\u000bi+\u000bi\u00001)\n\u0000t\nU\u0012(\u0000\u001b)[p\n2(\fi+\fi\u00001) +t\nU(2\u000b1\n2\ni+\u000b1\n2\ni+1+\u000b1\n2\ni\u00001)]\n+O(t2=U2): (37)\nIn the above derivation, we have used that \u0012(\u001b)\u0012(\u001b0) =\n\u0012(\u001b)\u000e\u001b;\u001b0. Notice that, sincet\nU\u001c1, in Eqs. (36) and (37)\nwe can identify the fields \u000b1\n2\ni\u0019aiA#and\u000bi\u0019(aiB1\"+\naiB2\")=p\n2, a result fully consistent with the low-energy spin\nconfiguration of the ferrimagnetic state discussed previously.\nAnalogously, for the high-energy bands, the opposite spin\nconfiguration is observed, with spin up (down) present at sites\nA(B1,B2).\nIntroducing Eqs. (36) and (37) into Eq. (24), with the aid\nof Eq. (25), we obtain a perturbative expression for H0(low-\nenergy sector) in terms of the spinless Grassmann fields up to\norderJ= 4t2=U:\nH0=\u0000JX\ni[\u000by\ni\u000bi+\u000b(1\n2)y\ni\u000b1\n2\ni\u0000\fy\ni\fi\u0000\f(1\n2)y\ni\f1\n2\ni]\n\u0000J\n2X\ni[\u000by\ni\u000bi+1+\u000b(1\n2)y\ni\u000b1\n2\ni+1\u0000\fy\ni\fi+1\u0000\f(1\n2)y\ni\f1\n2\ni+1+H.c:]\n+UX\ni[\fy\ni\fi+\f(1\n2)y\ni\f1\n2\ni+ey\ni#ei#]:(38)\nBy applying Fourier transform to the above expression and\nrearranging the terms, we obtain\nH0=\u0000X\nk2Jcos2(k=2)(\u000by\nk\u000bk+\u000b(1\n2)y\nk\u000b1\n2\nk)\n+X\nk[2Jcos2(k=2) +U](\fy\nk\fk+\f(1\n2)y\nk\f1\n2\nk)\n+U\n2X\nk\u001b(1\u0000\u001b)ey\nk\u001bek\u001b: (39)\nIn Fig. 1 we plot the electronic spectrum of the Hamilto-\nnianH0, both in the weak and strong-coupling regime: (a)\nEq. (29) for U= 2tand (b) Eqs. (29) and (39) for U= 12t\n(J= 4t2=U= 1=3), respectively, with t\u00111. We can notice\nthe presence of the shrinking phenomenon [7] as Uincreases\nfrom 2tto12t(strong-coupling regime) and that, for U= 12t,\nEq. (39) is a very good approximation to Eq. (29). Notice-\nably, thet\u001cUexpansion of the fields allow us to identify\n\u000b1\n2\nk\u0019akA#,\u000bk\u0019(akB1\"+akB2\")=p\n2(triplet state) and\nek\"\u0019(akB1\"\u0000akB2\")=p\n2(singlet state), as the low-energy\nspin configuration of the ferrimagnetic state with single occu-\npancy, where spins at sites A(B1;B2) are down (up), in agree-\nment with Lieb’s theorem [7, 8, 15].\nIn order to describe the most relevant low-energy pro-\ncesses that take place in this regime, one has to additionally\nproject out the high-energy bands from H0, that is, terms con-\ntaining only fields related to the high-energy bands are ex-\ncluded. Therefore, after the Legendre transformation, H0=\n\u0000P\ni;\u0011i@L0\n@(@\u001c\u0011i)@\u001c\u0011i+L0, where\u0011i=\u000bi;\u000b1\n2\ni, andei\"(fields\n-1 0 1\nk/π-2024 E(a)\nβkσ\nek↓\nek↑\nαkσU= 2t\n-1 0 1\nk/π-2061214(b)\nU= 12tβkβ1\n2\nk\nek↓\nek↑\nαkα1\n2\nkFigure 1. (Color online) Electronic spectrum of the Hamiltonian H0:\n(a) Eq. (29) for U= 2tand (b) Eqs. (29) and (39) for U= 12t(J=\n4t2=U= 1=3), witht\u00111. Notice the band shrinking phenomenon\nasUincreases from 2tto12t(strong-coupling regime). The t\u001c\nUexpansion of the fields identifies \u000b1\n2\nk\u0019akA#,\u000bk\u0019(akB1\"+\nakB2\")=p\n2andek\"\u0019(akB1\"\u0000akB2\")=p\n2, where spins at sites\nA(B1;B2) are down (up), in agreement with Lieb’s theorem [15].\nrelated to the low-energy bands), with@L0\n@(@\u001c\u0011i)=\u0011y\ni, the La-\ngrangian associated with H0(up to order J) is given by\nL0=X\ni;\u0011i\u0011y\ni@\u001c\u0011i\u0000JX\ni(\u000by\ni\u000bi+\u000b(1\n2)y\ni\u000b1\n2\ni)\n\u0000J\n2X\ni(\u000by\ni\u000bi+1+\u000b(1\n2)y\ni\u000b1\n2\ni+1+H.c:):(40)\nWe shall now focus on the U\u001dtperturbative expansion of\nLn, Eq. (22), which amounts to consider the most significant\nlow-energy processes, after the use of Eqs. (36) and (37) for\ndi\u001bandbi\u001bin terms of the spinless Grassmann fields. How-\never, terms allowing interband transitions between low- and\nhigh-energy bands do exist in Ln. In this context, we apply a\nsuitable second-order Rayleigh-Schr ¨odinger perturbation the-\nory [54, 55], consistent with the strong-coupling expansion,\nso that the modes associated with the high-energy bands are\neliminated. Lastly, by adding L0to the perturbative expansion\nofLn, which leads to the cancellation of the exchange terms\nin Eq. (40), the effective low-energy Lagrangian density of the\nAB2Hubbard model in the strong-coupling limit (up to order\nJ) reads:\nLeff(\u001c) =L(I)+L(II)+L(III)+L(IV); (41)\nwhere\nL(I)=X\ni\u000by\ni@\u001c\u000bi+X\ni\u000b(1\n2)y\ni@\u001c\u000b(1\n2)\ni+X\niey\ni\"@\u001cei\";(42a)7\nL(II)=X\ni\u001bn\n\u0012(\u0000\u001b)(U(b)y\ni@\u001cU(b)\ni)\u001b;\u001b\u000b(1\n2)y\ni\u000b(1\n2)\ni\n+\u0012(\u001b)1\n2[(U(d)y\ni@\u001cU(d)\ni)\u001b;\u001b+ (U(e)y\ni@\u001cU(e)\ni)\u001b;\u001b]\n\u0002(\u000by\ni\u000bi+ey\ni\"ei\") +\u0014\n\u0012(\u001b)1\n2[(U(d)y\ni@\u001cU(e)\ni)\u001b;\u001b\n+(U(e)y\ni@\u001cU(d)\ni)\u001b;\u001b]\u000by\niei\"+H.c.io\n; (42b)\nL(III)=\u0000tX\ni\u001bn\n\u0012(\u0000\u001b)(U(b)y\niU(d)\ni)\u001b;\u0000\u001b\u000b(1\n2)y\ni\u000bi\n+\u0012(\u001b)(U(d)y\niU(b)\ni+1)\u001b;\u0000\u001b\u000by\ni\u000b(1\n2)\ni+1\n+\u0012(\u0000\u001b)(U(b)y\niU(e)\ni)\u001b;\u0000\u001b\u000b(1\n2)y\niei\"\n+\u0012(\u001b)\u0010\nU(e)y\niU(b)\ni+1\u0011\n\u001b;\u0000\u001bey\ni\"\u000b(1\n2)\ni+1+H.c.\u001b\n;(42c)\nL(IV)=\u0000J\n4X\ni;i0=i;i+1;\u001b\u0012(\u001b)j(U(d)y\niU(b)\ni0)\u001b;\u001bj2\u000by\ni\u000bi\n\u0000J\n4X\ni;i0=i;i+1;\u001b\u0012(\u001b)j(U(e)y\niU(b)\ni0)\u001b;\u001bj2ey\ni\"ei\"\n\u0000J\n4X\ni;i0=i;i\u00001;\u001b\u0012(\u0000\u001b)[j(U(b)y\niU(d)\ni0)\u001b;\u001bj2\n+ (U(b)y\niU(e)\ni0)\u001b;\u001bj2]\u000b(1\n2)y\ni\u000b(1\n2)\ni; (42d)\nwhere\nU(b)\ni=UiA; U(d;e)\ni =1p\n2(UiB1\u0006UiB2); (43)\nin which case we took advantage of the symmetry of the AB2\nchain under the exchange operation B1$B2, in correspon-\ndence with Eq. (25). From the above equations, we see that the\nkinetic term is represented by L(I)and is related to the charge\ndegrees of freedom only, whereas L(II)describes the dynam-\nics of the spin degrees of freedom coupled to the charge fields.\nOn the other hand, L(III)exhibit first-neighbor hopping con-\ntributions between charge degrees of freedom in the presence\nofSU(2)gauge fields, while L(IV)is the spin exchange term\nin the presence of the charge Grassmann fields.\nIV . HALF-FILLING REGIME\nLet us now discuss some basic aspects of the localized mag-\nnetic properties related to the spin degrees of freedom. At half\nfilling, i.e.,\u000e= 0, we haveh\u000by\ni\u000bii= 1,h\u000b(1=2)y\ni\u000b(1=2)\nii= 1,\nhey\ni\"ei\"i= 1, andh\u000by\niei\"i= 0(no band hybridization) as the\nelectrons tend to fill up the lower-energy bands, whereas the\nhigher-energy ones remain empty. As a consequence, a fer-\nrimagnetic configuration of localized spins emerges, i.e., the\ncharge degrees of freedom are completely frozen, such that\nh\u000by\ni@\u001c\u000bii=h\u000b(1=2)y\ni@\u001c\u000b(1=2)\nii=hey\ni\"@\u001cei\"i= 0, with for-\nbidden hopping. Therefore, only terms from LIIandLIVinEqs. (42b) and (42d), respectively, give nonzero contributions\nand the resulting effective strong-coupling Lagrangian at half\nfilling, defined in Eq. (41), reads:\nLJ\neff=X\ni\u000b\u001b\u0012(pi\u000b\u001b)(Uy\ni\u000b@\u001cUi\u000b)\u001b;\u001b\n\u0000J\n4X\nhi\u000b;j\fi\u001b\u0012(pi\u000b\u001b)\f\f\f(Uy\ni\u000bUj\f)\u001b;\u001b\f\f\f2\n;(44)\nwhere the staggered factor pi\u000bwas defined in Eq. (11), and use\nwas made of the matrix transformations defined in Eq. (43) in\norder to sum up the squares of the SU(2)gauge field prod-\nucts in the exchange contribution from LIVin Eq. (42d).\nNow, using the following Legendre transform: HJ\neff =\n\u0000P\ni\u000b\u001b@LJ\neff\n@(@\u001cUi\u000b)\u001b;\u001b(@\u001cUi\u000b)\u001b;\u001b+LJ\neff;where@LJ\neff\n@(@\u001cUi\u000b)\u001b;\u001b=\n\u0012(pi\u000b\u001b)(Uy\ni\u000b)\u001b;\u001b;we get the respective quantum Heisenberg\nHamiltonian written in terms of the SU(2)gauge fields at half\nfilling as\nHJ\neff=\u0000J\n4X\nhi\u000b;j\fi\u001b\u0012(pi\u000b\u001b)\f\f\f(Uy\ni\u000bUj\f)\u001b;\u001b\f\f\f2\n: (45)\nFurther, using the definition of the SU(2)=U(1)unitary\nrotation matrix Eq. (14), it is possible to write [54–56]\f\f\f(Uy\ni\u000bUj\f)\u001b;\u001b\f\f\f2\n=1\n2(1 + ni\u000b\u0001nj\f);where ni\u000b=\nsin(\u0012i\u000b) [cos(\u001ei\u000b)^x+ sin(\u001ei\u000b)^y] + cos(\u0012i\u000b)^zis the unit vec-\ntor pointing along the local spin direction. Lastly, by using the\nconstraint as given in Eq. (19), we can identify the spin field\nfSi\u000bgat the single occupied sites:\nSi\u000b=pi\u000bni\u000b=2; (46)\nwherepi\u000b= +1 (\u00001)at sites\u000b=B1; B2(A), in order to\nobtain\nHJ\neff=JX\nih\n(SB1\ni+SB2\ni)\u0001(SA\ni+SA\ni+1)i\n\u0000JNc:(47)\nThe above expression is indeed that of the quantum antifer-\nromagnetic Heisenberg spin-1/2 model on the AB2chain in\nzero-field, which takes into account the effects of zero-point\nquantum spin fluctuations. In fact, to achieve this goal, we an-\nalyze the Hamiltonian, Eq. (47), by means of the spin-wave\ntheory, which has proved very successful in describing the\nproperties of the GS and low-lying excited states of spin mod-\nels. The predicted results provide a check of the consistency\nof our approach and will be fully used in our description of\nthe doped regime.\nWe shall first introduce boson creation and annihilation op-\nerators via the Holstein-Primakoff [58] transformation:\nSA;z\ni=\u0000S+ay\niai;\nSA;+\ni= (SA;\u0000\ni)y=p\n2Say\nifA(S);(48)\nfor a down-spin on the Asite, and\nSBl;z\ni=S\u0000by\nlibli;\nSBl;+\ni = (SBl;\u0000\ni)y=p\n2SfB(S)bli;(49)8\nfor an up-spin on the Blsite, withl= 1;2, and\nfr(S) =\u0010\n1\u0000nr\n2S\u00111=2\n= 1\u00001\n2nr\n2S+:::; (50)\nwhereSis the spin magnitude, and nr=ay\niaiorby\nlibli. The\noperatorsay\niandai(orby\nli,bli) satisfy the boson commutation\nrules. Under the above transformation, the spin Hamiltonian,\nEq. (47) is mapped onto the boson Hamiltonian:\nHJ\neff=E0\u0000JNc+H1+H2+O(S\u00001); (51)\nwhere\nE0=\u00004S2JNc; (52)\nis the classical GS energy and H1andH2are the quadratic\nand quartic (interacting) terms of the boson Hamiltonian, suit-\nable to describe the quantum AB2Heisenberg model via a\nperturbative series expansion in powers of 1=S. By Fourier\ntransforming the boson operators, we find\nH1= 2JSX\nk(2ay\nkak+X\nlby\nlkblk)\n+X\nk;l=1;22JS\rk(ay\nkby\nlk+akblk); (53)\nwhere we have defined the lattice structure factor as\n\rk=1\nzX\n\u001aeik\u001a= cos\u0012k\n2\u0013\n; (54)\nwithzdenoting the coordination number ( z= 4for theAB2\nchain), while \u001a=\u00061=2connects the nearest neighbors A-B1\nandA-B2linked sites of sublattices AandB, and\nH2=\u00003J\n2NX\n1234;l=1;2\u000e12;34n\n4\r1\u00004ay\n1a4by\nl3bl2\n+(\r1ay\n1by\nl4by\nl3bl2+\r1+2\u00003ay\n1ay\n2a3by\nl4+H.c.)o\n:(55)\nFor simplicity, we use the convention 1fork1,2fork2, and so\non. Also, the \u000e12;34=\u000e(k1+k2\u0000k3\u0000k4)is the Kronecker\n\u000efunction, and expresses the conservation of momentum to\nwithin a reciprocal-lattice vector G.\nWe shall consider H1first, which is the term leading to\nlinear spin-wave theory (LSWT). In fact, H1is diagonalized\nusing the following Bogoliubov transformation:\nak=uk\fk\u0000vk\u000by\nk;\nblk=1p\n2[uk\u000bk\u0000vk\fy\nk+ (\u00001)l\u0018k];withl= 1;2;(56)\n(uk;vk) =(3 +p\n9\u00008\r2\nk;2p\n2\rk)q\n(3 +p\n9\u00008\r2\nk)2\u00008\r2\nk; (57)\nwhereukandvksatisfy the constraint u2\nk\u0000v2\nk= 1. Thus,\nH1=E1+X\nk(\u000f0(\u000b)\nk\u000by\nk\u000bk+\u000f0(\f)\nk\fy\nk\fk+\u000f0(\u0018)\nk\u0018y\nk\u0018k);(58)E1=JSX\nk(q\n9\u00008\r2\nk\u00003); (59)\n\u000f0(\u000b;\f)\nk=JS(q\n9\u00008\r2\nk\u00071); \u000f0(\u0018)\nk= 2JS; (60)\nwhereE1is theO(S1)quantum correction to the GS energy,\nand\u000f0(\u000b;\f)\nk,\u000f0(\u0018)\nkare the three spin-wave branches provided\nby LSWT, both in agreement with previous results [19, 61].\nIn fact, it is well known that systems with a ferrimagnetic GS\nnaturally have ferromagnetic and antiferromagnetic spin-wave\nmodes as their elementary magnetic excitations (magnons).\nFor theAB2chain, there are three spin-wave branches: an\nantiferromagnetic mode ( \u000f0(\f)\nk) and two ferromagnetic ones\n(\u000f0(\u000b)\nkand\u000f0(\u0018)\nk). The mode \u000f0(\u000b)\nkis gapless at k= 0, i.e., the\nGoldstone mode, with a quadratic (ferromagnetic) dispersion\nrelation\u000f0(\u000b)\nk\u0018k2. The other two modes are gapped. No-\ntice that the gapped ferromagnetic mode \u000f0(\u0018)\nkis flat, and is\nclosely associated with ferrimagnetic properties at half fill-\ning [7, 17]. Since the dispersive modes preserve the local\ntriplet bond, they are identical to those found in the spin- 1=2\n- spin-1 chains [62–65]. These chains also exhibit interesting\nfield-induced Luttinger liquid behavior [66].\nNow, our aim is to obtain the leading corrections to LSWT,\ni.e., second-order spin-wave theory to the GS energy, sublat-\ntice magnetizations and Lieb GS total spin per unit cell. In do-\ning so, we develop a perturbative scheme for the description\nof this quartic term. First, we decompose the two-body terms\nby means of the Wick theorem, via normal-ordering protocol\nfor boson operators. Conservation of momentum to within a\nreciprocal-lattice vector, implies: k1=k+q,k2=p\u0000q,\nk3=kandk4=p. Then, we need to look at the possible\npairings of the 4 operators, as for example, in the first term of\nEq. (55):\nay\nk+qapby\nl;kbl;p\u0000q;ay\nk+qapby\nl;kbl;p\u0000q;ay\nk+qapby\nl;kbl;p\u0000q:\nUnder this procedure, and by substituting the Bogoliubov\ntransformation, Eqs. (56)-(57), into Eq. (55), we find\nH2=E2+X\nk(\u000e\u000f(\u000b)\nk\u000by\nk\u000bk+\u000e\u000f(\f)\nk\fy\nk\fk+\u000e\u000f(\u0018)\nk\u0018y\nk\u0018k);(61)\nwhere\nE2=Nc=\u00002J(q2\n1+q2\n2\u00003p\n2q1q2); (62)\nand the corresponding corrections for the spin-wave disper-\nsion relations read:\n\u000e\u000f(\u000b)\nk=J[u2\nk(p\n2q2\u00002q1) + 2v2\nk(p\n2q2\u0000q1)]\n+ 4J\rkukvk\u00143\n2p\n2q1\u0000q2\u0015\n+O(S\u00001);(63)\n\u000e\u000f(\f)\nkis obtained from \u000e\u000f(\u000b)\nkthrough the exchange of uk$\nvk, and\n\u000e\u000f(\u0018)\nk=J(p\n2q2\u00002q1) +O(S\u00001): (64)9\nIn Eqs. (62)-(64) above, the quantities q1andq2are defined\nby (thermodynamic limit)\nq1=1\n2\u0019\u0002\u0019\n\u0000\u0019dk(v2\nk); q 2=1\n2\u0019\u0002\u0019\n\u0000\u0019dk(\rkukvk):(65)\nWe remark that in deriving Eqs. (62)-(64), we have neglected\nterms containing anomalous products, such as, \u000by\nk\fy\nkand ver-\ntex corrections.\nLastly, the above results of our perturbative 1=Sseries ex-\npansion lead to the effective Hamiltonian:\nHJ\neff=EJ\nGS\u0000JNc+X\nk(\u000f\u000b\nk\u000by\nk\u000bk+\u000f(\f)\nk\fy\nk\fk+\u000f(\u0018)\nk\u0018y\nk\u0018k);\n(66)\nwhere\nEJ\nGS=E0+E1+E2; (67)\nwhich can be read from Eqs. (52), (59), and (62), respectively,\nis the second-order result up to O(1=S)for the GS energy,\nand\n\u000f(s)\nk=\u000f0(s)\nk+\u000e\u000f(s)\nk;withs=\u000b; \f; \u0018; (68)\nare the corresponding second-order spin-wave modes, where\nthe linear and the second-order correction terms are given by\nEq. (60) and Eqs. (63)-(65), respectively.\nA. Second-order spin-wave analysis\nOur perturbative 1=Sseries expansion approach is able to\nimprove the LSWT result for the gap \u0001 =Jof the antiferro-\nmagnetic mode, which should be compared with the second-\norder result derived from \u000f(\f)\nk, Eqs. (60), (63) and (68), at\nk= 0:\u0001 = (1 +p\n2q2)J'1:676J, in full agreement\nwith similar spin-wave calculations for AB2[19] and spin-\n1=2-spin- 1[64, 65] chains, and in agreement with numeri-\ncal estimates using exact diagonalization, \u0001 = 1:759J, for\nbothAB2[5] and spin- 1=2-spin- 1[63] chains. On the other\nhand, the LSWT predicts a gap \u0001flat=Jfor the flat fer-\nromagnetic mode ( \u000f(\u0018)\nk) inAB2chain, whereas our second-\norder spin-wave theory finds, using Eqs. (60), (64) and (68):\n\u0001flat= (1\u00002q1+p\n2q2)J'1:066J, in full agreement\nwith a similar spin-wave procedure [19]. Surprisingly, the es-\ntimated value from Exact Diagonalization (ED) [5]: \u0001flat=\n1:0004J, lies between these two theoretical values. In fact,\nanalytical approaches are still unable to reproduce the ob-\nserved level crossing found in numerical calculations [5, 19]\nfor the two ferromagnetic modes. This is probably due to the\nfact that the different symmetries exhibited by the localized\nexcitation (flat mode) and the ferromagnetic dispersive mode\nare not explicitly manifested in the analytical approaches, so\nthe levels avoid the crossing.B. Ground state energy\nIn the thermodynamic limit, the second-order result for the\nGS energy of the AB2chain per unit cell reads:\nEJ\nGS\nNc=\u00004JS2+JS\n2\u0019\u0002\u0019\n\u0000\u0019dk\u0012q\n9\u00008\r2\nk\u00003\u0013\n\u00002J(q2\n1+q2\n2\u00003p\n2q1q2): (69)\nWe remark that, at half filling, we shall not consider the con-\nstant term\u0000JNcin Eq. (51), with the purpose of compari-\nson with preceding results. Performing the integration over\nthe first BZ and taking S= 1=2, we obtain that the GS en-\nergy per site at zero-field is given by \u00000:4869J. This result\nagrees very well with values obtained using exact diagonal-\nization [45] (\u00000:485J) and DMRG [67] ( \u00000:4847J) tech-\nniques. For the spin- 1=2- spin-1 chain, the value obtained\nusing DMRG [62] is \u00000:72704J. To compare it with our find-\ning, we need to multiply this value by 2=3(ratio between the\nnumber of sites of the two chains), yielding \u00000:48469J.\nC. Sublattice magnetizations and Lieb GS total spin per unit\ncell\nIn order to derive results beyond LSWT, we intro-\nduce staggered magnetic fields coupled to spins SA;z\ni\nandSBl;z\ni, withl= 1;2, through the Zeeman terms:\n\u0000hAP\niSA;z\niand\u0000hBlP\niSBl;z\ni, which are added to HJ\neff\nin Eq. (47). Thus, hSA;ziandhSBl;zicorresponding\nto sublattices AandBlare obtained from hSA;zi=\n\u0000(1=Nc)P\ni=1;2[@Ei(hA)=@hA]jhA=0, and an analogous\nequation forhSBl;ziusing Eqs. (59) and (62):\n(hSA;zi;hSBl;zi) =\u0007S\u0006\u00121\n2;1\n4\u00131\n\u0019\u0002\u0019\n\u0000\u0019dkv2\nk\n\u0007\u00121\n2;1\n4\u0013q1\n\u0019S\u0002\u0019\n\u0000\u0019dk\r2\nk\n(9\u00008\r2\nk)3=2+O(1\nS2):(70)\nCarrying out the above integration, we obtain hSA;zi=\n\u00000:316343 andhSBl;zi= 0:408172 . These results are in\ngood agreement with those obtained using DMRG [11] and\nED [5] techniques: hSA;zi=\u00000:2925 andhSBl;zi=\n0:3962 , respectively, and with values for hSA;ziand\n2hSBl;zifor the spin- 1=2- spin-1 chain [62–65]. Although\nat zero temperature, the sublattice magnetizations are strongly\nreduced by quantum fluctuations, as compared with their clas-\nsical values, the unit cell magnetization remains SL\u00111=2,\nwhereSLis the Lieb GS total spin per unit cell, in full agree-\nment with Lieb’s theorem [5, 15] for bipartite lattices:\nSL=1\n2kNA\u0000NBk; (71)\nwithNA(NB)denoting the total number of spins in sublattice\nA(B)per unit cell.10\nV .t-JHAMILTONIAN: DOPING-INDUCED PHASES,\nGROUND STATE ENERGY AND TOTAL SPIN\nIn this section, we shall derive the corresponding t-J\nHamiltonian suitable to describe the strongly correlated AB2\nHubbard chain in the doped regime, in which case both charge\n(Grassmann fields) and spin [ SU(2)gauge fields] quantum\nfluctuations are considered on an equal footing. Indeed, the\nt-JHamiltonian can be derived by means of the following\nLegendre transformation to Eq. (41):\nHt-J\neff=\u0000X\ni;\u0016=b;d;e@Leff\n@(@\u001cU(\u0016)\ni)\u001b;\u001b(@\u001cU(\u0016)\ni)\u001b;\u001b\n\u0000X\ni;\u0017i@Leff\n@(@\u001c\u0017i)@\u001c\u0017i+Leff; (72)\nwhere@Leff\n@(@\u001c\u0017i)=\u0017y\niwith\u0017i=\u000bi;\u000b1\n2\ni;ei\";@Leff\n@(@\u001cU(b)\ni)\u001b;\u001b=\n\u0012(\u0000\u001b)(U(b)y\ni)\u001b;\u001b\u000b(1\n2)y\ni\u000b(1\n2)\ni, and@Leff\n@(@\u001cU(d;e)\ni)\u001b;\u001b=\n\u0012(\u001b)1\n2[(U(d;e)y\ni )\u001b;\u001b(\u000by\ni\u000bi+ey\ni\"ei\") + (U(e;d)y\ni )\u001b;\u001b(\u000by\niei\"+\ney\ni\"\u000bi)], from which we can write the effective t-JHamilto-\nnian as\nHt-J\neff=Ht+HJ; (73)\nwhere\nHt=\u0000tX\ni\u001bf\u0012(\u0000\u001b)(U(b)y\niU(d)\ni)\u001b;\u0000\u001b\u000b(1=2)y\ni\u000bi\n+\u0012(\u001b)(U(d)y\niU(b)\ni+1)\u001b;\u0000\u001b\u000by\ni\u000b(1=2)\ni+1\n+\u0012(\u0000\u001b)(U(b)y\niU(e)\ni)\u001b;\u0000\u001b\u000b(1=2)y\niei\"\n+\u0012(\u001b)(U(e)y\niU(b)\ni+1)\u001b;\u0000\u001bey\ni\"\u000b(1=2)\ni+1+H.c.g; (74)\nand\nHJ=\u0000J\n4X\ni;i0=i;i+1;\u001b\u0012(\u001b)j(U(d)y\niU(b)\ni0)\u001b;\u001bj2\u000by\ni\u000bi\n\u0000J\n4X\ni;i0=i;i+1;\u001b\u0012(\u001b)j(U(e)y\niU(b)\ni0)\u001b;\u001bj2ey\ni\"ei\"\n\u0000J\n4X\ni;i0=i;i\u00001;\u001b\u0012(\u0000\u001b)[j(U(b)y\niU(d)\ni0)\u001b;\u001bj2\n+j(U(b)y\niU(e)\ni0)\u001b;\u001bj2]\u000b(1\n2)y\ni\u000b(1\n2)\ni: (75)\nNotice that Eqs. (74) and (75) are identical to Eqs. (42c) and\n(42d), since Eqs. (42a) and (42b) were eliminated through the\nLegendre transformation.\nSome digression on Ht-J\neffis in order. One of the key prop-\nerties of quasi-1D interacting quantum systems is the phe-\nnomenon of spin-charge separation, leading to the formation\nof spin and charge-density waves, which move independently\nand with different velocities. It has been demonstrated [24]\nthat for\u000e > 2=3the low-energy physics of the doped AB2\nHubbard chain in the U=1coupling limit is described interms of the Luttinger-liquid model, with the spin and charge\ndegrees of freedom decoupled. Most importantly, it has been\nshown that for the AB2t-JHubbard chains [25] charge and\nspin quantum fluctuations are practically decoupled, as sug-\ngested by the emergence of charge-density waves in anti-\nphase with the modulation of the ferrimagnetic order. One\ncan make use of this feature to formally split each term of the\nt-JHamiltonian, Eq. (73)-(75) , into a product of two inde-\npendent terms acting on different Hilbert spaces, i.e., we can\nenforce spin-charge separation and calculate the charge and\nspin correlation functions in a decoupled fashion.\nTherefore, from the above discussion, we shall consider\nthat the charge correlation functions are well described by an\neffective spinless tight-binding model [24, 55, 68], since the\nhole (charge) density waves develop along the x-axis and in\nanti-phase with the modulation of the ferrimagnetic structure,\nas numerically observed in Fig. 2(b) of Ref. [25]. So, using\nEqs. (33), with a=2!a(effective lattice spacing of the lin-\near chain: distance between AandBsites, see Fig. 2(a) of\nRef. [25]), we find\nh\u000b(1=2)y\ni\u000bii=1\nNcX\nkk0e\u0000ik(xi\u00001)eik0xih\t0j\u000by\nk\u000bk0j\t0i\n=1\n\u0019\u0002kF(\u000e)\n\u0000kF(\u000e)eikdk=2\n\u0019sin[kF(\u000e)]; (76)\nwithj\t0ibeing the hole-doped ferrimagnetic GS,\nwherekF(\u000e) =\u0019Nh\nN\u0011\u0019\u000e is the Fermi wave vec-\ntor of the spinless tight-binding holes. In the\nsame fashion: h\u000by\ni\u000b(1=2)\ni+1i=2\n\u0019sin[kF(\u000e)] and\nh\u000b(1=2)y\niei\"i=hey\ni\"\u000b(1=2)\ni+1i= 0; whileh\u000by\ni\u000bii =\nh\u000b(1=2)y\ni\u000b(1=2)\nii=hey\ni\"ei\"i= (1\u00001\n2\u0019\u0001kF(\u000e)\n\u0000kF(\u000e)dk) =\n(1\u0000\u000e). Here, we remark that the itinerant holes away from\nhalf filling are associated with the lower-energy dispersive \u000bk\nand\u000b(1=2)\nkbands [see Fig. 1(b) in Sec. (II)], thus contributing\nto the kinetic Hamiltonian in Eq. (74). On the other hand,\nthe local correlations related to the lower-energy bands\n\u000bk,\u000b(1=2)\nk, andek\", contribute equally to the exchange\nHamiltonian in Eq. (75). Thereby, using the above tight-\nbinding results for the charge correlation functions, Ht-J\neff\nin Eqs. (73)-(75) gives rise to the \u000e-dependent Hamiltonian,\nHt-J\neff(\u000e) =Ht\neff(\u000e) +HJ\neff(\u000e), written below:\nHt-J\neff(\u000e) =\u0000t2\n\u0019sin[kF(\u000e)]X\ni[(U(b)y\niU(d)\ni)#\"\n+ (U(d)y\niU(b)\ni+1)\"#+H.c.]\n\u0000J(1\u0000\u000e)\n4X\nhi\u000b;j\fi\u001b\u0012(pi\u000b\u001b)j(Uy\ni\u000bUj\f)\u001b;\u001bj2;(77)\nwhere the sum over \u001bwas evaluated in Eq. (74) and the square\nof theSU(2)gauge field products in the exchange contribu-\ntion have been summed up in Eq. (75), so that this contribu-\ntion is just (1\u0000\u000e)timesHJ\neffat half filling, Eq. (45), or\nalternatively, in terms of spin fields, Eq. (47), or spin-waves,\nEqs. (66)-(68). On the other hand, the SU(2)gauge fields ma-11\ntrix elements:\u0010\nU(b)y\niU(d)\ni\u0011\n#\"and\u0010\nU(d)y\niU(b)\ni+1\u0011\n\"#, that ap- pear in the kinetic contribution of Eq. (77), can be written in\nterms of the spin fields [55, 56] as\n(U(b)y\niU(d)\ni)#\"+H.c.=X\nl1p\n2\u0012q\n1\u00002SBl;z\ni\u00002SA;z\ni+ 4SA;z\niSBl;z\ni+q\n1 + 2SA;z\ni+ 2SBl;z\ni+ 4SA;z\niSBl;z\ni\u0013\n;(78)\nand\u0010\nU(d)y\niU(b)\ni+1\u0011\n\"#is obtained from Eq. (78) through the replacement SA;z\ni!SA;z\ni+1, in which case we took advantage of the\nU(1)gauge freedom and Eq. (46). Notice that these square-root matrix elements depend on z-spin components only.\nAt this stage, it will prove useful, in the calculation of the GS total spin in the doped regime, to consider Ht-J\neff(\u000e;h)which\ndescribes the system in the presence of a homogeneous magnetic field h=h^ z= (\u0000hA+hB1+hB2)^ z, where the staggered fields\npoint along the local corresponding magnetizations in the ferrimagnetic phase have the same magnitude h. The magnetic field\ncouple with the spin fields through the Zeeman term (see Sec. IV C) andwith the charge degrees of freedom through the magnetic\norbital coupling in the Landau gauge: A=hx^ y. Since our aim is to study doping effect on the magnetization, we shall assume\nvanishingly small magnetic field in the context of linear response theory and perturbative expansion in the strong-coupling\nregime. Additionally, the magnetic orbital coupling can be considered through the so-called Peierls substitution [27, 69]: t!\ntei\u0001j\f\ni\u000bA\u0001dl, wherei\u000bandj\fare first-neighbor sites, and the flux quantum \u001e0=hc=e\u00111. If one consider that the carrier is\nat the siteiA, we have four hopping possibilities: iA!iB1;2andiA!(i+ 1)B1;2, so the total phase \u001eacquired by the\ncarrier in this prescription satisfies Stokes’ theorem: \u001e=\u0017\nunit cellA\u0001dl=s\nSh\u0001dS=ha2(a\u00111). We also remark that,\nin order to obtain real values for the zero-field staggered magnetizations, we have considered, for convenience, an imaginary\ngauge transformation [56, 70]: A!iA. Therefore, by placing Eq. (78) and the similar matrix element into the kinetic term in\nEq. (77), making the above Peierls substitution, and using the Holstein-Primakoff and Bogoliubov transformations introduced in\nEqs. (48)-(50) and Eqs. (56)-(57), respectively, up to order O(S\u00001), we arrive at the following diagonalized kinetic Hamiltonian\nHt\neff(\u000e;h):\nHt\neff(\u000e;h) =\u00004p\n2\n\u0019te\u0000(\u0000hA+hB1+hB2)sin[kF(\u000e)]X\nk[4S\u00003v2\nk\u0000(u2\nk+ 2v2\nk)\u000by\nk\u000bk\u0000(2u2\nk+v2\nk)\fy\nk\fk\u0000\u0018y\nk\u0018k];(79)\nwhere the doped-induced contributions for the spin dispersion relations are evidenced in the last three terms. On the other hand,\nby adding the Zeeman terms (see Sec. IV C) to the exchange contribution HJ\neff(\u000e), given in Eq. (77), we obtain HJ\neff(\u000e;h).\nLastly, by adding the kinetic and the exchange contributions, we arrive at the effective t-JHamiltonian in the presence of a\nmagnetic field:\nHt-J\neff(\u000e;h) =\u00004p\n2\n\u0019te\u0000(\u0000hA+hB1+hB2)sin[kF(\u000e)]X\nk(4S\u00003v2\nk) +J(1\u0000\u000e)(EJ\nGS\u0000JNc)\n+X\nk[\u000f(\u000b)\nk(\u000e)\u000by\nk\u000bk+\u000f(\f)\nk(\u000e)\fy\nk\fk+\u000f(\u0018)\nk(\u000e)\u0018y\nk\u0018k]\u0000hAX\niSA;z\ni\u0000hB1X\niSB1;z\ni\u0000hB2X\niSB2;z\ni;(80)\nwhereEJ\nGSis given by Eq. (67) and (69), and the corresponding spin-wave modes [see Eqs. (79), (60), (63)-(65), and (68)] of\nthe dopedAB2t-Jchain read:\n\u000f(\u000b)\nk(\u000e) =4p\n2\n\u0019tsin(\u0019\u000e)[u2\nk+ 2v2\nk] + (1\u0000\u000e)(\u000f0(\u000b)\nk+\u000e\u000f(\u000b)\nk); (81)\n\u000f(\f)\nk(\u000e)is obtained from \u000f(\u000b)\nk(\u000e)through the exchange uk$vkand the replacement \u000b!\f, while\n\u000f(\u0018)\nk(\u000e) =4p\n2\n\u0019tsin(\u0019\u000e) + (1\u0000\u000e)(\u000f0(\u0018)\nk+\u000e\u000f(\u0018)\nk): (82)\nWe find it instructive to comment on the analytical structure of the above equations. Firstly, we mention the presence of the\nBogoliubov parameters [see Eqs. (57)] in a symmetric form in the kinetic terms of Eq. (81) and its analogous for \u000f(\f)\nk(\u000e); besides,\nalthough the flat mode is strongly affected by the presence of holes, it remains dispersionless. In addition, using Eqs. (80) and\n(65), the total GS energy (no spin-wave excitations) per unit cell in the thermodynamic limit is readily obtained:\nEt-J\nGS(\u000e;h)=Nc=\u00004p\n2\n\u0019te\u0000(\u0000hA+hB1+hB2)sin(\u0019\u000e)(4S\u00003q1)\n+ (1\u0000\u000e)(EJ\nGS=Nc\u0000J)\u0000hSA;zihA\u0000X\nl=1;2hSBl;zihBl; (83)12\nwherehSA;ziandhSBl;ziare the calculated sublattice magnetizations, at half filling and zero-field, given by Eqs. (70).\nIn subsections V A, V B, and V C, we will show that the\nunderlying competing physical mechanisms: the magnetic\norbital response and the Zeeman contribution embedded in\nEqs. (80)-(83) will dramatically affect the behavior of the sys-\ntem under hole doping and, in particular, will lead to spiral\nIC spin structures, the breakdown of the spiral ferrimagnetic\nGS at a critical value of the hole doping, a region of phase\nseparation, and RVB states at \u000e\u00191=3.\nA. Doped regime: Spin-wave modes\nBefore we go one step further to discuss relevant macro-\nscopic quantities, i.e., the GS energy and total spin in the\ndoped regime, we shall first undertake a detailed study, at a\nmicroscopic level, of the hole-doping effect on the calculated\nspin-wave branches given by Eqs. (81)-(82).\nFig. 2 depicts the second-order spin-wave dispersion rela-\ntions atJ=t= 0:3and for the indicated values of \u000e. Without\nloss of generality, we set t= 1in our numerical computations.\nAt half filling, the antiferromagnetic mode \u000f(\f)\nk, together with\nthe two ferromagnetic modes: the dispersive \u000f(\u000b)\nkand the flat\none\u000f(\u0018)\nk, are shown in Fig. 2(a), which are defined in Eq. (68),\nand can be plotted using Eqs. (60) and (63)-(65).\nAs the hole doping increases slightly, the abrupt decrease\nof the peaks at k= 0 andk=\u0019of the numerical DRMG\nstructure factor (see Fig. 3 of Ref. [25]), associated with the\nferrimagnetic order, manifests itself here through the opening\nof a gap in the ferromagnetic Goldstone mode \u000f(\u000b)\nk, as seen in\nFig. 2(b), thus indicating that the system loses its long-range\norder. Note that the antiferromagnetic mode \u000f(\f)\nkis also sim-\nilarly shifted. On the other hand, although the dispersion re-\nlation is modified for small values of the wave vector k, the\nminimum value of \u000f(\u000b)\nkstill remains at k= 0 up to the onset\nof the formation of spiral IC spin structures at \u000ec(IC) = 0:043\n(a value that should be compared with the numerical DMRG\nestimate of\u000e\u00190:055\u00060:012), characterized by the flatten-\ning of the dispersive spin-wave branches around zero. Upon\nfurther increase of \u000e, two minima form (around k= 0) and\nmove away from each other as one enhances the hole doping.\nThis behavior is the signature of the occurrence of spiral IC\nspin structures (see Fig. 3 of Ref. [25]).\nFig. 2(c) shows the onset of phase separation (PS) at\n\u000e(PS) = 0:165forJ= 0:3, which is characterized by the\noverlap of the two ferromagnetic modes at k= 0. The sig-\nnature of this regime is the spatial coexistence of two phases:\nspiral IC spin structures at \u000e(PS) = 0:165and RVB states\nat\u000e\u00191=3, in very good agreement with the numerical es-\ntimate of\u000eIC\u0000PS\u00190:16[25]. At\u000e\u00191=3, the flat mode\nhas the lowest energy, as illustrated in Fig. 2(d). This be-\nhavior indicates that the RVB state is the stable phase at\n\u000e\u00191=3andJ= 0:3[25], and also in agreement with the\nnumerical DMRG studies [11, 24] and analytical prediction at\nU=1[55].\nOnset of IC\nOnset of PS\n(b) (a)\n(c) (d)Figure 2. (Color online) Evolution of the zero-field second-order\nspin-wave dispersion relations of the AB2t-Jchain as a function of\nhole doping ( \u000e): dispersive ferromagnetic \u000f(\u000b)\nkand antiferromagnetic\n\u000f(\f)\nkmodes and the flat ferromagnetic one \u000f(\u0018)\nk, at (a) half filling; (b)\nthe onset of the spiral IC spin structures at \u000ec(IC) = 0:043, in which\ncase the flattening of the gap of \u000f(\u000b)\nkaroundk= 0 is observed; (c)\nthe onset of PS at \u000e(PS) = 0:165, characterized by the overlap of\nthe two ferromagnetic modes at k= 0and by the spatial coexistence\nof two phases: spiral IC spin structures, with modulation fixed at\n\u000e(PS) , and RVB states at \u000e\u00191=3. (d) At\u000e= 1=3the flat mode\npresents the lowest energy, thus indicating that the short-range RVB\nstate is the stable phase.\nIn order to better understand the rich variety of doping-\ninduced phases in the system, in Fig. 3 we plot the evolution of\nthe wave vector kmincorresponding to the local minimum of\n\u000f(\u000b)\nk(\u000e), upon increasing the hole doping \u000efrom 0to1=3. The\nwave vector kminremains zero until it hits the onset doping\nvalue\u000ec(IC) = 0:043, beyond which a square-root growth be-\nhavior takes place [71]: [\u000e\u0000\u000ec(IC)]1=2(blue line), for \u000eclose\nto\u000ec(IC). The square-root growth behavior is the signature\nof the occurrence of a second-order quantum phase transition\nfrom the doped ferrimagnetic phase to the IC spiral ferrimag-\nnetic state with a non-zero value of the total GS spin, SGS.\nThis result is supported by the behavior of \u0001k\u0011kmax\u0000\u0019\nat which the local maximum of the numeric DMRG structure\nfactorS(k)neark=\u0019is observed, as shown in the inset of\nFig. 3 (taken from the inset of Fig. 3(b) of Ref. [25]). For\nfurther increase of hole doping our result deviates from the\nsquare-root growth behavior and some very interesting fea-\ntures are to be noticed. The value of \u000ec= 0:08indicates\nthe breakdown of the total SGSin the IC phase, as will be\nconfirmed by the explicit calculation of SGS, a macroscopic\nquantity, in Section V C. Thus, for 0:08<\u000e < 0:165the sys-\ntem displays an IC phase with zero SGS, in agreement with\nthe DMRG data (see Fig. 1(c) of Ref. [25]). At \u000e(PS) = 0:16513\n0 0.043 0.08 0.165 1/300.10.20.3kmin/π\nDoped FerriIC\nSGS/negationslash= 0IC\nSGS= 0PS\nδc(IC)δcδδ(PS)∼\nRVB\nAnalytic\n[δ−δc(IC)]1/2\n0 0.05 0.1δ00.10.20.30.4\n∆k/π\nFigure 3. (Color online) Evolution of kmin(value ofkat the local\nminimum of \u000f(\u000b)\nk(\u000e)neark= 0) as a function of \u000e: doped ferri-\nmagnetism for 0< \u000e < \u000ec(IC)\u00190:043; spiral IC spin structures\nwith non-zero (zero) SGSfor\u000ec(IC)< \u000e < \u000ec\u00190:08(\u000ec< \u000e <\n\u000e(PS)\u00190:165), with a second-order quantum phase transition at\n\u000ec(IC) characterized by a square-root behavior [\u000e\u0000\u000ec(IC)]1=2(blue\nline), and a first-order transition at \u000e(PS) involving the IC spin struc-\nture, with modulation fixed at \u000e(PS) , and short-range RVB states at\nhole concentration 1/3. The inset shows DMRG data from Ref. [25]\nfor\u0001k\u0011kmax\u0000\u0019as a function of \u000e, wherekmaxis the value of\nkat the local maximum of the structure factor S(k)neark=\u0019, in\nqualitative agreement with the second-order transition at \u000ec(IC) .\nthe system exhibits a first-order transition accompanied by the\nspatial phase separation regime: the IC phase with zero SGS\nand modulation fixed by \u000e(PS) in coexistence with the short-\nrange RVB states at \u000e\u00191=3, also consistent with the DMRG\ndata plotted in Fig. 4 of Ref. [25].\nLastly, we emphasize that, despite the occurrence of several\ndoping-induced phases in the DMRG studies [25]: Lieb fer-\nrimagnetism, spiral IC spin structures, RVB states with finite\nspin gap, phase separation, and Luttinger-liquid behavior, it\nis surprising and very interesting that the second-order spin-\nwave modes remain stable up to \u000e\u00191=3, with predictions\nin very good agreement with the DMRG studies [25]. In this\ncontext, it is worth mentioning the long time studied case of\nrare earth metals [72], where an external magnetic field can in-\nduce non-trivial phase transitions involving spiral spin struc-\ntures, well described by spin-wave theory.\nB. Doped regime: Ground state energy\nPerforming the integration over the first BZ in Eqs. (65) and\n(69) and setting S= 1=2in Eq. (83), we find that the AB2t-J\nground state energy per unit cell as a function of hole doping\nin zero-field reads:\nEt-J\nGS(\u000e)=JNc=\u00001:9543t\nJsin (\u0019\u000e)\u00002:4608 (1\u0000\u000e):(84)\nWe shall now examine the case of small hole doping away\nfrom half filling, i.e., with hole concentration ranging from\n\u000e= 0 up to\u000e= 0:2for two values of J:0:1and0:3. In\nFig. 4, we show the evolution of the GS energy per unit cell\n0 0.05 0.1 0.15 0.2\nδ0\n-5\n-10\n-15Et-J\nGS(δ)/JNc\nAnalyticJ= 0.3\nNumericJ= 0.3\nAnalyticJ= 0.1\nNumericJ= 0.1Figure 4. (Color online) Analytical prediction for the GS energy per\nunit cell of the AB2t-Jchain as a function of doping, and compar-\nison with numerical data from DMRG technique for J=t= 0:1and\nJ=t= 0:3[25]. At half filling ( \u000e= 0), both results meet at the\nexpected prediction [25]: \u0019\u00002:4678 . Note that we have added the\nterm\u0000JNcwith the intention of comparison with numerical calcu-\nlation.\nof theAB2t-Jmodel as a function of hole doping for both\nmentioned values of J, and the comparison was made with the\nnumerical DMRG data [25]. From the two results at J= 0:3,\nthe only quantitative difference induced by the increase of the\nhole concentration is a crossing feature around \u000e\u00190:1, where\nour analytical result slightly change its behavior by lowering\nthe energy with respect to the numerical data [25]. In fact, be-\ncause our model assumes a ferrimagnetic state as the starting\npoint, this change of behavior suggests that we have entered in\na region of strong magnetic instabilities, and possibly indicat-\ning a smooth transition to an incommensurate phase with zero\nGS total spin beyond \u000e\u00190:1, as confirmed by the numerical\ndata in Ref. [25] and illustrated in Fig. 3. On the other hand,\natJ= 0:1, although our results reproduce the numerical data\nwith an acceptable agreement, we observe a discrepancy that\nincreases with \u000e. The cause of such discrepancy will be dis-\ncussed in the next subsection.\nWith the purpose of determining the interplay between the\ncontribution of magnetic exchange and the itinerant kinetic\nenergy to the zero-field GS energy Eq. (83), we take J= 0:3\nand show its evolution with doping in Fig. 5. We can see\nin the insets, Fig. 5(a) and Fig. 5(b), the competitive behav-\nior of the two energetic contributions, i.e., the contribution of\nthe exchange energy increases linearly with \u000e, while a practi-\ncally linear decrease of the hopping term is observed as one\nenhances the hole doping. This competition indicates that a\nphase transition to a paramagnetic phase should occur at some\ncritical concentration value.\nC. Doped regime: Ground state total spin\nThe existence of a transition from an IC spiral ferrimag-\nnetic phase to an IC paramagnetic one is a most interest-14\n0 0.05 0.1 0.15 0.2\nδ0\n-2\n-4\n-6Et-J\nGS(δ)/JNc\nJ= 0.3\nExchange\nHopping0 0.05 0.1 0.15 0.2\nδ-2\n-2.2\n-2.4EJ(δ)/JNc\n0 0.05 0.1 0.15 0.2\nδ0\n-1\n-2\n-3Et(δ)/JNc\nFigure 5. (Color online) Ground-state energy per unit cell for the\nAB2t-Jchain as a function of \u000eforJ= 0:3. In the insets, we\nillustrate the two energetic contribution due to (a) exchange and (b)\nhopping terms.\ning feature observed numerically in doped AB2t-JHubbard\nchains [25]. In order to firmly corroborate the mentioned\ntransition, we have calculated the GS total spin per unit cell\nas a function of hole doping, SGS(\u000e) =P\n\u000bhS\u000b;zi(\u000e),\nwith\u000b=A;B 1;B2, by means of the zero-field derivative\nof Eq. (83):\nhS\u000b;zi(\u000e) =\u0000(1=Nc)[@Et-J\nGS(\u000e;h)=@h\u000b]jh\u000b=0:(85)\nWe thus find\nhS\u000b;zi(\u000e) =hS\u000b;zi\u00064p\n2\n\u0019tsin(\u0019\u000e)(4S\u00003q1);(86)\nwhere +(\u0000) corresponds to sublattice \u000b=A(\u000b=B1;2),\nandhSA;ziandhSBl;ziare given by Eqs. (70). Therefore,\nby performing the integration over the first BZ of the three\ncontributions in Eq. (86), we finally obtain:\nSGS(\u000e)\nSL= 1\u00003:9086 sin(\u0019\u000e); (87)\nwhereSL=P\n\u000bhS\u000b;zi= 1=2is Lieb’s reference value for\nthe GS total spin per unit cell at half filling and zero-field (see\nSection IV C).\nIn Fig. 6 we plot the evolution of SGS, normalized by SL,\nas a function of \u000e, and compare it with the numerical data\nfrom DMRG and Lanczos techniques [25], for J= 0:3(red\nsquares) and J= 0:1(blue circles). In the latter (former) case,\nthe system undergoes a transition from the modulated itin-\nerant ferrimagnetic phase to an incommensurate phase with\nzero (nonzero) SGS. Notice that, in both cases, the transition\nis characterized by a decrease of SGSfromSLto0or to a\nresidual value, regardless of the value that SGStakes after the\ntransition. Indeed, at J= 0:1and\u000e > 0:1, the formation\nof magnetic polarons (onset of the Nagaoka phenomena that\nsets in asU!1 ) with charge-density waves in phase with\nthe modulation of the ferrimagnetic structure, as indicated by\nFigure 6. (Color online) Ground-state total spin SGSper unit cell\n(solid magenta line), normalized by its value in the undoped regime:\nSL=1\n2, as a function of hole doping \u000efor the indicated values of\nJ. In the figure, \u000ec\u00190:08indicates the critical value of doping\nat which the magnetic order is suppressed and a second-order phase\ntransition takes place.\nthe DMRG data [25], leads to an incommensurate phase with\nnonzeroSGS.\nMost importantly, we can observe in Fig. 6 that the value of\nSGSdecreases practically linearly with \u000euntil the magnetic\norder is completely suppressed at \u000ec\u00190:08. This behavior is\nsupported by numerical results [25], particularly in the regime\nwhere the Nagaoka phenomenon is not manifested, that is, at\nJ= 0:3, as indicated in Fig. 3. In this regime, spin and charge\nquantum fluctuations destabilize the ferrimagnetic structure\nand trigger a transition to an incommensurate paramagnetic\nphase at\u000ec, withSGS\u0018(\u000e\u0000\u000ec)!0.\nVI. CONCLUSIONS\nIn summary, we have presented a detailed analytical study\nof the large-U Hubbard model on the quasi-one-dimensional\nAB2chain. We used a functional integral approach combined\nwith a perturbative expansion in the strong-coupling regime\nthat allowed us to properly analyze the referred system at and\naway from half filling.\nAt half filling, our model was mapped onto the quantum\nHeisenberg model, and analyzed through a spin-wave pertur-\nbative series expansion in powers of 1=S. We have demon-\nstrated that the GS energy, spin-wave modes, and sublat-\ntice magnetizations are in very good agreement with previ-\nous results. In the challenging hole doping regime away\nfrom half filling, the corresponding t-J(= 4t2=U)Hamilto-\nnian was derived. Further, under the assumption that charge\nand spin quantum correlations are decoupled, the evolution\nof the second-order spin-wave modes in the doped regime\nhas unveiled the occurrence of spatially modulated spin struc-\ntures and the emergence of phase separation (first-order tran-\nsition) in the presence of resonating-valence-bond states. The\ndoping-dependent GS energy and total spin per unit cell are15\nalso calculated, in which case the collapse of the spiral mag-\nnetic order at a critical hole concentration was observed.\nRemarkably, our above-mentioned analytical results in the\ndoped regime are in very good agreement with density matrix\nrenormalization group studies, where our assumption of spin-\ncharge decoupling is numerically supported by the formation\nof charge-density waves in anti-phase with the modulation of\nthe ferrimagnetic structure.\nFinally, we stress that our reported results evidenced that\nthe present approach, also used in a study on the compatibil-\nity between numerical and analytical outcomes of the large-U\nHubbard model on the honeycomb lattice, was proved suitablefor theAB2chain (a quasi-1D system), where the impact of\ncharge and spin quantum fluctuations are expected to manifest\nin a stronger way. We thus conclude that our approach offers a\nquite powerful analytical description of hole-doping induced\nphases away from half filling in low-dimensional strongly-\ncorrelated electron systems.\nACKNOWLEDGMENTS\nWe appreciate interesting discussions with R. R.\nMontenegro-Filho. This work was supported by CNPq,\nCAPES and FACEPE/PRONEX (Brazilian agencies).\n[1] S. Sachdev, Quantum Phase Transitions , 2nd ed. (Cambridge\nUniversity Press, 2011).\n[2] M. A. Continentino, Quantum scaling in many-body systems ,\n2nd ed. (Cambridge University Press, 2017).\n[3] M. D. Coutinho-Filho, R. R. Montenegro-Filho, E. P. Raposo,\nC. Vitoriano, and M. H. Oliveira, J. Braz. Chem. Soc. 19, 232\n(2008).\n[4] N. B. Ivanov, Condensed Matter Phys 12(2009).\n[5] R. R. Montenegro-Filho and M. D. Coutinho-Filho, Physica A\n357, 173 (2005).\n[6] J. Silvestre and R. Hoffmann, Inorg. Chem. 24, 4108 (1985).\n[7] A. M. S. Mac ˆedo, M. C. dos Santos, M. D. Coutinho-Filho, and\nC. A. Mac ˆedo, Phys. Rev. Lett. 74, 1851 (1995).\n[8] G.-S. Tian and T.-H. Lin, Phys. Rev. B 53, 8196 (1996).\n[9] F. C. Alcaraz and A. L. Malvezzi, J. Phys. A 30, 767 (1997).\n[10] E. P. Raposo and M. D. Coutinho-Filho, Phys. Rev. Lett. 78,\n4853 (1997); Phys. Rev. B 59, 14384 (1999).\n[11] G. Sierra, M. A. Mart ´ın-Delgado, S. R. White, D. J. Scalapino,\nand J. Dukelsky, Phys. Rev. B 59, 7973 (1999).\n[12] M. A. Mart ´ın-Delgado, J. Rodriguez-Laguna, and G. Sierra,\nPhys. Rev. B 72, 104435 (2005).\n[13] E. Lieb and D. Mattis, J. Math. Phys. 3, 749 (1962).\n[14] E. H. Lieb and F. Y . Wu, Phys. Rev. Lett. 20, 1445 (1968).\n[15] E. H. Lieb, Phys. Rev. Lett. 62, 1201 (1989).\n[16] E. H. Lieb, in The Hubbard Model: Its Physics and Math-\nematical Physics, Nato ASI, Series B: Physics ,edited by D.\nBaeriswyl, D. K. Campbell, J. M. P Carmelo, F. Guinea and E.\nLouis, V ol. 343 (Plenum, New York, 1995).\n[17] H. Tasaki, J. Phys. Condens. Matter 10, 4353 (1998); Prog.\nTheor. Phys. 99, 489 (1998).\n[18] G.-S. Tian, J. Stat. Phys. 116, 629 (2004).\n[19] S. Yamamoto and J. Ohara, Phys. Rev. B 76, 014409 (2007).\n[20] R. R. Montenegro-Filho and M. D. Coutinho-Filho, Phys. Rev.\nB78, 014418 (2008).\n[21] K. Hida and K. Takano, Phys. Rev. B 78, 064407 (2008).\n[22] T. Shimokawa and H. Nakano, Journal of the Physical Society\nof Japan 81, 084710 (2012).\n[23] S. C. Furuya and T. Giamarchi, Phys. Rev. B 89, 205131 (2014).\n[24] R. R. Montenegro-Filho and M. D. Coutinho-Filho, Phys. Rev.\nB74, 125117 (2006).\n[25] R. R. Montenegro-Filho and M. D. Coutinho-Filho, Phys. Rev.\nB90, 115123 (2014).\n[26] T. Giamarchi, Quantum Physics in One Dimension (Clarendon;\nOxford University Press, 2004).\n[27] F. H. L. Essler, H. Frahm, F. G ¨ohmann, A. Kl ¨umper, andV . E. Korepin, The one-dimensional Hubbard model (Cam-\nbridge University Press, 2005).\n[28] J. M. P. Carmelo and P. D. Sacramento, Phys. Rep. 749(2018).\n[29] F. D. M. Haldane, Phys. Rev. Lett. 67, 937 (1991); Y .-S. Wu,\nPhys. Rev. Lett. 73, 922 (1994).\n[30] C. Vitoriano, R. R. Montenegro-Filho, and M. D. Coutinho-\nFilho, Phys. Rev. B 98, 085130 (2018).\n[31] A. A. Lopes and R. G. Dias, Phys. Rev. B 84, 085124 (2011);\nA. A. Lopes, B. A. Z. Ant ´onio, and R. G. Dias, Phys. Rev. B\n89, 235418 (2014).\n[32] M. Drillon, M. Belaiche, P. Legoll, J. Aride, A. Boukhari, and\nA. Moqine, J. Magn. Magn. Mater 128, 83 (1993).\n[33] A. A. Belik, A. Matsuo, M. Azuma, K. Kindo, and M. Takano,\nJ. Solid State Chem. 178, 709 (2005).\n[34] M. Matsuda, K. Kakurai, A. A. Belik, M. Azuma, M. Takano,\nand M. Fujita, Phys. Rev. B 71, 144411 (2005).\n[35] W. Guo, Z. He, and S. Zhang, J. Alloys Compd. 717, 14 (2017).\n[36] M. Verdaguer, A. Gleizes, J. P. Renard, and J. Seiden, Phys.\nRev. B 29, 5144 (1984).\n[37] A. S. F. Ten ´orio, R. R. Montenegro-Filho, and M. D. Coutinho-\nFilho, J. Phys. Condens. Matter 23, 506003 (2011).\n[38] J. Stre ˇcka and T. Verkholyak, J. Low. Temp. Phys , 1 (2016);\nSee also: X. Yan, Z.-G. Zhu, and G. Su, AIP Advances 5,\n077183 (2015).\n[39] H. Kikuchi, Y . Fujii, M. Chiba, S. Mitsudo, T. Idehara, T. Tone-\ngawa, K. Okamoto, T. Sakai, T. Kuwai, and H. Ohta, Phys.\nRev. Lett. 94, 227201 (2005).\n[40] M. Oshikawa, M. Yamanaka, and I. Affleck, Phys. Rev. Lett.\n78, 1984 (1997).\n[41] M. Hase, M. Kohno, H. Kitazawa, N. Tsujii, O. Suzuki,\nK. Ozawa, G. Kido, M. Imai, and X. Hu, Phys. Rev. B 73,\n104419 (2006).\n[42] K. C. Rule, A. U. B. Wolter, S. S ¨ullow, D. A. Tennant, A. Br ¨uhl,\nS. K ¨ohler, B. Wolf, M. Lang, and J. Schreuer, Phys. Rev. Lett.\n100, 117202 (2008).\n[43] F. Aimo, S. Kr ¨amer, M. Klanj ˇsek, M. Horvati ´c, C. Berthier, and\nH. Kikuchi, Phys. Rev. Lett. 102, 127205 (2009).\n[44] H. Jeschke, I. Opahle, H. Kandpal, R. Valent ´ı, H. Das, T. Saha-\nDasgupta, O. Janson, H. Rosner, A. Br ¨uhl, B. Wolf, M. Lang,\nJ. Richter, S. Hu, X. Wang, R. Peters, T. Pruschke, and A. Ho-\nnecker, Phys. Rev. Lett. 106, 217201 (2011).\n[45] K. Takano, K. Kubo, and H. Sakamoto, J. Phys. Condens. Mat-\nter8, 6405 (1996).\n[46] K. Okamoto, T. Tonegawa, Y . Takahashi, and M. Kaburagi, J.\nPhys. Condens. Matter 11, 10485 (1999).16\n[47] K. Okamoto, T. Tonegawa, and M. Kaburagi, J. Phys. Condens.\nMatter 15, 5979 (2003).\n[48] A. S. F. Ten ´orio, R. R. Montenegro-Filho, and M. D. Coutinho-\nFilho, Phys. Rev. B 80, 054409 (2009).\n[49] K. Morita, M. Fujihala, H. Koorikawa, T. Sugimoto, S. Sota,\nS. Mitsuda, and T. Tohyama, Phys. Rev. B 95, 184412 (2017).\n[50] M. Fujihala, H. Koorikawa, S. Mitsuda, K. Morita, T. Tohyama,\nK. Tomiyasu, A. Koda, H. Okabe, S. Itoh, T. Yokoo, S. Ibuka,\nM. Tadokoro, M. Itoh, H. Sagayama, R. Kumai, and Y . Mu-\nrakami, Sci. Rep. 7, 16785 (2017).\n[51] E. Dagotto, Science 309, 257 (2005).\n[52] T.-R. T. Han, F. Zhou, C. D. Malliakas, P. M. Duxbury, S. D.\nMahanti, M. G. Kanatzidis, and C.-Y . Ruan, Sci. Adv. 1(2015),\n10.1126/sciadv.1400173.\n[53] E. Dagotto, Rep. Prog. Phys. 62, 1525 (1999).\n[54] Z. Y . Weng, D. N. Sheng, C. S. Ting, and Z. B. Su, Phys. Rev.\nLett. 67, 3318 (1991); Phys. Rev. B 45, 7850 (1992).\n[55] M. H. Oliveira, E. P. Raposo, and M. D. Coutinho-Filho, Phys.\nRev. B 80, 205119 (2009).\n[56] F. G. Ribeiro and M. D. Coutinho-Filho, Phys. Rev. B 92,\n045105 (2015).\n[57] Z.-C. Gu, H.-C. Jiang, D. N. Sheng, H. Yao, L. Balents, and\nX.-G. Wen, Phys. Rev. B 88, 155112 (2013).\n[58] E. Fradkin, Field theories of condensed matter physics , 2nd ed.\n(Cambridge University Press, 2013).\n[59] J. W. Negele and H. Orland, Quantum many-particle systems ,Frontiers in physics (Addison-Wesley Pub. Co., 1988).\n[60] W. K. Tung, Group theory in physics (World Scientific, New\nYork, 1985).\n[61] C. Vitoriano, F. B. D. Brito, E. P. Raposo, and M. D. Coutinho-\nFilho, Mol. Cryst. Liq. Cryst 374, 185 (2002).\n[62] S. K. Pati, S. Ramasesha, and D. Sen, Phys. Rev. B 55, 8894\n(1997).\n[63] S. Yamamoto, S. Brehmer, and H.-J. Mikeska, Phys. Rev. B 57,\n13610 (1998).\n[64] N. B. Ivanov, Phys. Rev. B 62, 3271 (2000).\n[65] S. Yamamoto, Phys. Rev. B 69, 064426 (2004).\n[66] W. M. da Silva and R. R. Montenegro-Filho, Phys. Rev. B 96,\n214419 (2017).\n[67] H. Niggemann, G. Uimin, and J. Zittartz, J. Phys. Condens.\nMatter 9, 9031 (1997).\n[68] C. Vitoriano and M. D. Coutinho-Filho, Phys. Rev. B 82,\n125126 (2010).\n[69] J. Vidal, B. Douc ¸ot, R. Mosseri, and P. Butaud, Phys. Rev. Lett.\n85, 3906 (2000); Z. Gul ´acsi, A. Kampf, and D. V ollhardt, Phys.\nRev. Lett. 99, 026404 (2007).\n[70] N. Hatano and D. R. Nelson, Phys. Rev. Lett. 77, 570 (1996).\n[71] S. Chakrabarty, V . Dobrosavljevi ´c, A. Seidel, and Z. Nussinov,\nPhys. Rev. E 86, 041132 (2012).\n[72] B. R. Cooper and R. J. Elliott, Phys. Rev. 131, 1043 (1963); See\nalso: A. R. Mackintosh and H. B. Møller, Spin Waves, inMag-\nnetic Properties of Rare Earth Metals, edited by R. J. Elliott\n(Plenum Press, 1972), Chapter 1, p. 187." }, { "title": "1901.03072v2.Ultrafast_magnetization_dynamics_in_uniaxial_ferrimagnets_with_compensation_point__GdFeCo.pdf", "content": "Ultrafast magnetization dynamics in uniaxial ferrimagnets with compensation point.\nGdFeCo\nM. D. Davydova,1, 2,\u0003K. A. Zvezdin,1, 2A. V. Kimel,3, 4and A. K. Zvezdin1, 2,y\n1Prokhorov General Physics Institute of the Russian Academy of Sciences, 119991 Moscow, Russia\n2Moscow Institute of Physics and Technology (State University), 141700 Dolgoprudny, Russia\n3Moscow Technological University (MIREA), 119454 Moscow, Russia\n4Institute for Molecules and Materials, Radboud University, Nijmegen 6525 AJ, The Netherlands\nWe derive an e\u000bective Lagrangian in the quasi-antiferromagnetic approximation that allows to\ndescribe the magnetization dynamics for uniaxial f-d(rare-earth - transition metal) ferrimagnet\nnear the magnetization compensation point in the presence of external magnetic \feld. We perform\ncalculations for the parameters of GdFeCo, a metallic ferrimagnet with compensation point that is\none of the most promising materials in ultrafast magnetism. Using the developed approach, we \fnd\nthe torque that acts on the magnetization due to ultrafast demagnetization pulse that can be caused\neither by ultrashort laser or electrical current pulse. We show that the torque is non-zero only in\nthe non-collinear magnetic phase that can be acquired by applying external magnetic \feld to the\nmaterial. The coherent response of magnetization dynamics amplitude and its timescale exhibits\ncritical behavior near certain values of the magnetic \feld corresponding to a spin-\rop like phase\ntransition. Understanding the underlying mechanisms for these e\u000bects opens the way to e\u000ecient\ncontrol of the amplitude and the timescales of the spin dynamics, which is one of the central problems\nin the \feld of ultrafast magnetism.\nPACS numbers: 79.20.Ds, 75.50.Gg, 75.30.Kz, 75.78.-n\nINTRODUCTION\nMost of the prominent advances in the \feld of ultrafast magnetism have been achieved by using thermal mechanism\nof magnetization control [1{7]. These studies rooted from the pioneering work by Beaurepaire et al. [8] on ultrafast\nlaser-induced demagnetization of Ni. In this experiment, partial destruction of magnetic order was found at much\nfaster rates that were believed to be possible prior to that publication. Since then, the \feld of ultrafast magnetism has\nbeen rapidly growing and the possible channels of ultrafast angular momentum transfer have been studied extensively\n[9]. Ultrafast demagnetization can be achieved by applying ultrashort laser pulses [1{7], or, alternatively, by using\nshort pulses of electric currents [10, 11].\nIn the last decades, GdFeCo and other rare-earth - transition metal compounds (RE-TM) have been in the center\nof attention in this regard [12]. For example, all-optical switching has been demonstrated for the \frst time in GdFeCo\n[13]. It was found that the switching is possible due to di\u000berent rates of sublattice demagnetization, which enables\nultrafast magnetization reversal to occur because of the angular momentum conservation [1, 2].\nIn many RE-TM compounds, GdFeCo and TbFeCo being part of them, realization of the magnetization compen-\nsation point is possible. At this point, the magnetizations of the two antiferromagnetically coupled sublattices with\ndi\u000berent dependencies on temperature become equal and the total magnetization of the material turns to zero. In\nthe presence of the external magnetic \feld, a record-breaking fast subpicosecond magnetization switching was found\nin GdFeCo across the compensation point [13]. In addition, a number of anomalies in the magnetic response was\nobserved near this point [14{16], which has never been explained theoretically. All said above illustrates the impor-\ntance of understanding the role of the compensation point in the dynamics and working out an appropriate tool for\nits description.\nE\u000ecient control of the amplitude and the timescales of the response of the magnetic system to an ultrafast de-\nmagnetizing impact on a medium is one of the most important issues in the area of ultrafast magnetism nowadays\n[9, 14, 15]. Understanding of the mechanisms and of the exhaustive description of the subsequent spin dynamics is\nalso a long-standing goal that will help to promote the achievements of this area towards practical applications in\nmagnetic recording [3, 17], magnonics [18] and spintronics [19]. In this work, we expand the understanding of response\nof magnetic system of a uniaxial f-dferrimagnet near the compensation point in the external magnetic \feld to an\nultrafast demagnetizing pulse, which can be induced either by a femtosecond laser or an electric current pulse. We\npresent a theoretical model and calculations, which allow to describe the ultrafast response of the system that resides\nin an angular phase before the impact. We show that in this magnetic phase the coherent precessional response is\npossible and the subsequent magnetization dynamics may become greatly nonlinear and is governed by large inter-\nsublattice exchange \feld [20]. We derive the e\u000bective Lagrangian that governs the dynamics of the system near thearXiv:1901.03072v2 [cond-mat.mtrl-sci] 28 Jan 20192\ncompensation point and obtain the torque acting on the magnetizations of the two sublattices due to demagnetiza-\ntion. In ref. [15], the critical response of the amplitude and the time of the signal rise have been found in GdFeCo\nin external magnetic \feld along the easy axis. At given laser pump \ruences, the response was found to be negligible\nin collinear phase, but it was dramatically large in angular one. We elaborate on this example and show that the\ncritical behavior of the response is the consequence of the second-order magnetic phase transition from collinear to an\nangular in the external magnetic \feld. We \fnd that these e\u000bects are pronounced in the vicinity of the compensation\npoint, where the phase transitions cross each other[21{23]. Thus, the proposed model explains a range of important\nexperimental observations as well as allows for developments of methods and tools of magnetization control by setting\nthe temperature near the compensation point and applying magnetic \feld. Moreover, by changing the composition\nof the alloy, the [24], the position of the magnetization compensation point can be tuned arbitrary close to the room\ntemperature. Our results might open new ways for technologies for ultrafast optical magnetic memory.\nEFFECTIVE LAGRANGIAN AND RAYLEIGH DISSPATION FUNCTION\nOur approach is based on Landau-Lifshitz-Gilbert equations for a two-sublattice (RE-TM) ferrimagnet. These\nequations are equivalent to the following e\u000bective Lagrangian and Rayleigh dissipation functions:\nL=Mf\n\r(1\u0000cos\u0012f)@'f\n@t+Md\n\r(1\u0000cos\u0012d)@'d\n@t\u0000\b(Mf;Md;H); (1)\nR=Rf+Rd;Rf;d=\u000bMf;d\n2\r\u0010\n_\u00122\nf;d+ sin2\u0012f;d_'2\nf;d\u0011\n(2)\nwhere\ris the gyromagnetic ratio, MdandMfare the magnetizations, \u0012d(TM) and\u0012f(RE) are the polar, 'dand\n'fare the azimuthal angles of d- andf- sublattices correspondingly in the spherical system of coordinates with z-axis\naligned along the external magnetic \feld H. \b(Mf;Md;H) is the thermodynamic potential for the system that we\ntake in the following form:\n\b =\u0000MdH+\u0015MdMf\u0000MfH\u0000Kf(Mfn)2\nM2\nf\u0000Kd(Mdn)2\nM2\nd; (3)\nwhere\u0015is the intersublattice exchange constant, nis the direction of the easy axis and Kf;dare the anisotropy\nconstants for f- andd- sublattices, respectively.\nNext, we transfer to description in terms of the antiferromagnetic L=MR\u0000Mdand the total magnetization\nM=MR+Mdvectors. In the vicinity of the compensation point the di\u000berence between the sublattice magnetizations\njMR\u0000Mdj\u001cLis small. The two vectors are parametrized using the sets of angles \u0012;\"and';\f, which are de\fned\nas:\n\u0012f=\u0012\u0000\"; \u0012d=\u0019\u0000\u0012\u0000\";\n'f='+\f; 'd=\u0019+'\u0000\f:(4)\nIn this case the antiferromagnetic vector is naturally de\fned as L= (Lsin\u0012cos';Lsin\u0012sin';Lcos\u0012).\nFollowing the work [25] we use the quasi-antiferromagnetic approximation to describe the dynamics near the mag-\nnetization compensation point. F In this approximation the canting angles are small \"\u001c1,\f\u001c1, and we can\nexpand the Lagrangian (1) and the corresponding thermodynamic potential up to quadratic terms in small variables:\nL=\u0000m\n\r_'cos\u0012\u0000M\n\rsin\u0012\u0010\n_'\"\u0000\f_\u0012\u0011\n\u0000\b;\n\b =\u0000K(l;n)2\u0000Hmcos\u0012\u0000\"MH sin\u0012+\u000e\n2\u0000\n\"2+ sin2\u0012\f2\u0001\n:(5)\nHerem=MR\u0000Md,M=MR+Md,K=KR+Kdis the e\u000bective uniaxial anisotropy constant, l=L=Lis\nthe unit antiferromagnetic vector \u000e= 2\u0015MdMRand we assume the anisotropy to be weak K\u001c\u0015M. For GdFeCo\nwith 24% Gd and compensation point near 283 K, we assume the following values of parameters: M\u0019800 emu/cc,\nK= 1:5\u0002105erg/cc,\u0015= 18:5 T/\u0016B,\u000e\u0019109erg/cc and mchanges in the range between 150 emu/cc and \u000050\nemu/cc at \felds H\u0019H\u0003\u00194 T. The characteristic values of small angles \u000fand\fare of the order of 10\u00002.3\nNext, we exclude the variables \"and\fby solving the Euler-Lagrange equations. Substituting them into the\nLagrangian (5), we obtain the e\u000bective Lagrangian, which describes the dynamics of a uniaxial ferrimagnet in the\nvicinity of the compensation point:\nLeff=\u001f?\n2 _\u0012\n\r!2\n+mcos\u0012\u0012\nH\u0000_'\n\r\u0013\n+\u001f?\n2sin2\u0012\u0012\nH\u0000_'\n\r\u00132\n+K(l;n)2; (6)\n\beff(H) =\u0000mHcos\u0012\u0000\u001f?\n2H2sin2\u0012\u0000K(l;n)2; (7)\nReff=\u000bM\n2\r\u0010\n_\u00122+ sin2\u0012_'2\u0011\n(8)\nwhere\u001f=2M2\n\u000e. In GdFeCo \u001f\u00191:6\u000210\u00003and\u000b\u00190:05. In the derivation above we assumed the gyrotropic factor\n\rand Gilbert damping constant \u000bto be the equal for both sublattices. Taking into account the di\u000berence between\nthese values for di\u000berent sublattices will lead to the angular momentum compensation e\u000bect at certain temperature.\nThe Lagrangian, Rayleigh function and equations of motion preserve the same form in this case if we substitute the\nparameters \rand\u000bwith temperature-dependent factors e\rande\u000bde\fned as:\n1\ne\r=1\n\u0016\r\u0012\n1 +M\nm\rf\u0000\rd\n\rf+\rd\u0013\n=Md\n\rd\u0000Mf\n\rf\n(Md\u0000Mf);1\n\u0016\r=1\n2\u00121\n\rd+1\n\rf\u0013\n;e\u000b=(\u000bd\rf+\u000bf\rd)\n(\rf+\rd)1\n1 +M\nm\rf\u0000\rd\n\rf+\rd(9)\nThis allows to reproduce the angular moment compensation phenomenon, which was studied experimentally in ref.\n[14].\nEXCITATION OF THE SPIN DYNAMICS\nThe proposed approach presents a powerful tool allowing analyzing coherent magnetization dynamics in ferrimagnets\nthat occurs under a broad range of conditions. Let us consider the following example that poses an important problem\nin the \feld of ultrafast magnetism. An femtosecond laser pulse strikes the uniaxial ferrimagnet (for instance, of\nGdFeCo, TbFeCo type) in the presence of external static magnetic \feld. The impact of the laser pulse leads to the\ndemagnetization of one or both of the sublattices. What coherent magnetization dynamics will occur as a consequence\nof this impact? The proposed model can be further developed in order to answer to this question and is applicable\nfor small values of demagnetization \u000eM.\nIn our framework the spin dynamics in ferrimagnet is described by Euler-Lagrange equations of the formd\ndt@L\n@_q\u0000@L\n@q=\n\u0000@R\n@_q, whereq=\u0012; ' are the polar and azimuthal angles describing the orientation of the antiferromagnetic vector\nL, correspondingly. Let us consider a particular case when the easy magnetization axis is aligned with the external\nmagnetic \feld, which leads to the presence of azimuthal symmetry in the system. In this case n= (0;0;1). In this\nparticular case the Euler-Lagrange equations can be rewritten as:\n\u001f?\n\r2\u0012=@Leff\n@\u0012\u0000@Reff\n@_\u0012;d\ndt@Leff\n@_'=\u0000@Reff\n@_'(10)\nThe nonlinear equations that are similar to Eqs. (10) and describe the spin dynamics of two-sublattice ferrimagnets\nwere obtained in the work [26] under the conditions H= 0 andReff= 0. Over the short time of demagnetization the\nsecond equation can be approximately treated as a conservation law and the conserving quantity (angular momentum\nof magnetization precession J) stays approximately constant as @L=@'= 0 due to the Noether theorem:\nJ=@Leff\n@_'=\u00001\n\r\u0014\nmcos\u0012+\u001f?sin2\u0012\u0012\nH\u0000_'\n\r\u0013\u0015\n=const (11)\nLet the moment of time t= 0\u0000denote the moment before the laser pulse impact and system initially is in the\nground state de\fned by the ground state angles \u0012(0\u0000) =\u00120,'(0\u0000) ='0, and their derivatives _ '(0\u0000) = 0, _\u0012(0\u0000) = 0.4\nDepending on the external parameters and preparation of the sample, the system might reside in one of the two\npossible antiferromagnetic collinear phases or in angular phase, which are separated by the magnetic phase transition\nlines [27]. If the demagnetization due to the laser pulse action is small, it produces the changes in the values of Mf,Md\nandMof the order of percent or less, whereas the change of m(which is approximately equal to total magnetization\nnear the compensation point) may be of several orders of magnitude, as its value is almost compensated. In what\nfollows, we assume that the demagnetization is associated only with change of m, namelym=m0+ \u0001m(t). As we\nwill see below, the change in this quantity already leads to several drastic e\u000bect in dynamics.\nTherefore, the conservation law (11) leads to the emergence of azimuthal dynamics _ '(t) at the demagnetization\ntimescales (\u0001 t) due to demagnetization pulse \u0001 m(t):\n_'(t) =\r\u0001m(t)\n\u001f?cos\u00120\nsin2\u00120(12)\nWe see that the torque is non-zero only in the angular phase, where 0 <\u00120<\u0019. Emergence of the azimuthal spin\nprecession as a result of demagnetization of the medium is similar to the well-known Einstein-de-Haas e\u000bect, where\nthe demagnetization leads to azimuthal precession of the body. Subsequently, this azimuthal spin dynamics leads to\nthe emergence of polar dynamics \u0012(t), which is most commonly measured in pump-probe experiments of ultrafast\nmagnetism, by acting as an e\u000bective \feld Heff=H\u0000_'\n\rin the Lagrangian (5). We can then view the Lagrangian\nas depending only on variable \u0012and the e\u000bective \feld Heff. At demagnetization \u000em\u00180:01Min GdFeCo the value\nof _'can reach up to 1 THz, and the corresponding e\u000bective magnetic \feld is of the order of 10 T. Note that initial\nstate of the system corresponds to the condition@\b\n@\u00120(Heff=H) = 0. We can rewrite the Euler-Lagrange equation\nfrom eq. (10) for polar angle as follows:\n\u001f?\n\r2\u0012+@\b(Heff)\n@\u0012=\u0000\u000bM\n\r_\u0012: (13)\nOr, alternatively:\n\u001f?\n\r2\u0012+msin\u0012Heff\u0000\u001f?sin\u0012cos\u0012\u0012\nH2\neff\u00002K\n\u001f?\u0013\n=\u0000\u000bM\n\r_\u0012 (14)\nBy integrating this equation over the short demagnetization pulse duration \u0001 twe obtain the state of system after\nthe laser pulse impact at t= 0+, which is characterized by the initial conditions\n\u0012(0+) =\u00120;_'(0+); '(0+) =Z\u0001t\n0_'(t)dt;_\u0012(0+) =Z\u0001t\n0\u0012(t)dt\nThe value \u0001 tis of the order of the optical pulse length. It may also include the time of restoration of the magnetization\nlength (or the value of m). After the moment of time 0+ free magnetization precession occurs in the model. Analysis\nof the spin dynamics under laser pump excitation will lead to emergence of critical dynamics near the second-order\nphase transitions to the collinear phases where \u0012= 0;\u0019, as is already seen from (12). We will discuss this behavior\nbelow.\nCRITICAL DYNAMICS\nIn a simple case of a quick decay of demagnetization (at the exciton relaxation timescales) with \u0001 m(t) = \u0001m,\n0\u0001t, we obtain the initial condition from (14):\n_\u0012(+0)\u0019\u0014\n\u0000\u0012\n2cos2\u00120\nsin\u00120+ sin\u00120\u0013\nH+m0\n\u001f?cos\u00120\nsin\u00120+\u0001m\n\u001f?cos\u00120\nsin3\u00120\u0015\r2\n\u001f?\u0001m\u0001t=B(\u00120)\u0001m+O(\u0001m2): (15)\nThis quantity de\fnes the initial angular momentum of the polar spin precession that is induced in the system due to\nthe optical spin torque created by the femtosecond laser pulse. The amplitude of oscillations is proportional to the\ninitial condition (15). Its dependence on the external magnetic \feld is illustrated in Fig. 1 for di\u000berent temperatures\nfor magnetic parameters of GdFeCo uniaxial ferrimagnet. At low values of external magnetic \felds there is only\ncollinear ground state in the ferrimagnet and above certain \feld Hsfthe transition to an angular state occurs [21].\nThe schematic of the magnetic phase diagram for GdFeCo is shown in insertions in Fig. 1. At T= 275 K and T= 2885\nFIG. 1. The amplitude of the magnetization precessional response after the demagnetization due to the femtosecond laser\npulse action in GdFeCo ferrimagnet near the compensation point at di\u000berent temperatures. Insertions: the schematic of the\nmagnetic phase diagram. There are two antiferromagnetic collinear phases with Mddirected along(opposite) to the external\nmagnetic \feld above(below) the compensation temperature TM. They are separated by the \frst-order phase transition line\n(blue). Above them, an area where the angular phase exists, which is \flled with gray color. The black solid lines are the\nsecond-order phase transition lines. The dashed lines corresponds to the \fxed temperature in the plot. The red dot is the point\nof phase transition for this temperature.\nK the phase transitions are of the second order, which corresponds to a smooth transition from angle \u00120= 0 to\u00120>0,\nand the divergence of the response occurs at Hsf. Immediately above the compensation temperature the transition is\nof the \frst order and the behavior of the response above the is more complex; however, there is no critical divergence.\nThe critical behavior of the signal amplitude was observed experimentally for GdFeCo in ref. [15].\nAnother feature in the dynamics described by the proposed model is the critical behavior of the characteristic\ntimescales that occurs in the vicinity of the second-order phase transitions. To demonstrate this e\u000bect analytically,\nwe assume small deviations of \u0012during oscillations: \u0012(t) =\u00120+\u000e\u0012(t). We obtain:\n\u000e\u0012+!2\nr(\u00120)\u000e\u0012=\u0000\u000b!ex\u000e_\u0012; (16)\nwhere!2\nr(\u00120) =\r2h\nm\n\u001f?Hcos\u00120+\u0010\n2K\n\u001f?\u0000H2\u0011\ncos 2\u00120i\n,!ex=\rM\n\u001f?. The initial conditions are \u0012(0) =\u00120and eq.\n(15). In the limit of small oscillations and !r<\u000b!ex=2 (is ful\flled near the second-order transition) the solution has\nthe form\u000e\u0012(t) =Ae\u0000\ftsinh!t, where\f=\u000b!ex=2,!2=\f2\u0000!2\nr,A=B(\u00120)=!. The rise time can be estimated from\nthe condition _\u0012(\u001crise) = 0:\n\u001crise\u0019atanh!\n\f\n!=atanhp\n\f2\u0000!2r\n\fp\n\f2\u0000!2r(17)\nThe time of the oscillations decay (relaxation time) is proportional to the imaginary part of eigenfrequency and can\nbe estimated by the following expression:\n\u001crelax\u00194\u0019\f\n!2r: (18)\nNear second-order phase transition the mode softening occurs and the eigenfrequency turns to zero: !r!0,\nand we observe growth of the both timescales. The critical behavior of the rise time has been observed in GdFeCo\nexperimentally [15] and the typical values of \u001crisewere of the order of 10 ps.6\nCONCLUSIONS\nTo sum up, the developed theoretical model based on quasi-antiferromagnetic Lagrangian formalism proved to\nbe suitable for description of the coherent ultrafast response of RE-TM ferrimagnets near the compensation point\ndue to an ultrashort pulse of demagnetization in the presence of external magnetic \feld. We have found that the\ntorque acting on magnetizations is non-zero in the noncollinear phase only. We have explained the experimentally\nobserved critical behavior of the response amplitude and characteristic timescales as the consequence of the second-\norder magnetic phase transition from collinear to an angular in the external magnetic \feld and the mode softening\nnear it. These e\u000bects are vivid in the vicinity of the compensation point in external magnetic \feld. Understanding\nthe ultrafast response to demagnetizing optical or electrical pulses and subsequent spin dynamics can facilitate future\ndevelopments in the \felds of ultrafast energy-e\u000ecient magnetic recording, magnonics and spintronics.\nACKNOWLEDGMENTS\nThis research has been supported by RSF grant No. 17-12-01333.\n\u0003davydova@phystech.edu\nyzvezdin@gmail.com\n[1] T. Ostler, J. Barker, R. Evans, R. Chantrell, U. Atxitia, O. Chubykalo-Fesenko, S. El Moussaoui, L. Le Guyader, E. Men-\ngotti, L. Heyderman, et al. , Nature communications 3(2012), 10.1038/ncomms1666.\n[2] I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius, H. D urr, T. Ostler, J. Barker, R. Evans, R. Chantrell, et al. ,\nNature 472, 205 (2011).\n[3] C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk, A. Tsukamoto, A. Itoh, and T. Rasing, Phys. Rev. Lett. 99, 047601\n(2007).\n[4] S. Wienholdt, D. Hinzke, K. Carva, P. M. Oppeneer, and U. Nowak, Phys. Rev. B 88, 020406 (2013).\n[5] A. R. Khorsand, M. Savoini, A. Kirilyuk, A. V. Kimel, A. Tsukamoto, A. Itoh, and T. Rasing, Phys. Rev. Lett. 108,\n127205 (2012).\n[6] R. Moreno, T. A. Ostler, R. W. Chantrell, and O. Chubykalo-Fesenko, Phys. Rev. B 96, 014409 (2017).\n[7] U. Atxitia, T. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell, and O. Chubykalo-Fesenko, Phys. Rev. B 87, 224417\n(2013).\n[8] E. Beaurepaire, J.-C. Merle, A. Daunois, and J.-Y. Bigot, Phys. Rev. Lett. 76, 4250 (1996).\n[9] A. Kirilyuk, A. V. Kimel, and T. Rasing, Rev. Mod. Phys. 82, 2731 (2010).\n[10] R. B. Wilson, J. Gorchon, Y. Yang, C.-H. Lambert, S. Salahuddin, and J. Bokor, Phys. Rev. B 95, 180409 (2017).\n[11] Y. Yang, R. B. Wilson, J. Gorchon, C.-H. Lambert, S. Salahuddin, and J. Bokor, Science Advances 3(2017), 10.1126/sci-\nadv.1603117, http://advances.sciencemag.org/content/3/11/e1603117.full.pdf.\n[12] R. Chimata, L. Isaeva, K. K\u0013 adas, A. Bergman, B. Sanyal, J. H. Mentink, M. I. Katsnelson, T. Rasing, A. Kirilyuk,\nA. Kimel, O. Eriksson, and M. Pereiro, Phys. Rev. B 92, 094411 (2015).\n[13] C. D. Stanciu, A. Tsukamoto, A. V. Kimel, F. Hansteen, A. Kirilyuk, A. Itoh, and T. Rasing, Phys. Rev. Lett. 99, 217204\n(2007).\n[14] C. D. Stanciu, A. V. Kimel, F. Hansteen, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing, Phys. Rev. B 73, 220402\n(2006).\n[15] J. Becker, A. Tsukamoto, A. Kirilyuk, J. C. Maan, T. Rasing, P. C. M. Christianen, and A. V. Kimel, Phys. Rev. Lett.\n118, 117203 (2017).\n[16] Z. Chen, R. Gao, Z. Wang, C. Xu, D. Chen, and T. Lai, Journal of Applied Physics 108, 023902 (2010),\nhttps://doi.org/10.1063/1.3462429.\n[17] A. Moser, K. Takano, D. T. Margulies, M. Albrecht, Y. Sonobe, Y. Ikeda, S. Sun, and E. E. Fullerton, Journal of Physics\nD: Applied Physics 35, R157 (2002).\n[18] B. Lenk, H. Ulrichs, F. Garbs, and M. Mnzenberg, Physics Reports 507, 107 (2011).\n[19] J. Walowski and M. Mnzenberg, Journal of Applied Physics 120, 140901 (2016), https://doi.org/10.1063/1.4958846.\n[20] M. Liebs, K. Hummler, and M. F ahnle, Phys. Rev. B 46, 11201 (1992).\n[21] B. Goransky and A. Zvezdin, JETP Lett. 10, 196 (1969).\n[22] A. Zvezdin and V. Matveev, JETP 35, 140 (1972).\n[23] C. K. Sabdenov, M. Davydova, K. Zvezdin, D. Gorbunov, I. Tereshina, A. Andreev, and A. Zvezdin, J. Low. Temp. Phys\n43, 551 (2017).\n[24] M. Ding and S. J. Poon, Journal of Magnetism and Magnetic Materials 339, 51 (2013).\n[25] A. Zvezdin, JETP Lett. 28, 553 (1979); arXiv preprint arXiv:1703.01502 (2017).\n[26] B. Ivanov and A. Sukstanskii, Zhurnal Eksperimental'noi i Teoreticheskoi Fiziki 84, 370 (1983).7\n[27] A. Zvezdin, Handbook of Magnetic Materials 9, 405 (1995)." }, { "title": "2205.13802v3.Magnonic_Casimir_Effect_in_Ferrimagnets.pdf", "content": "Magnonic Casimir Effect in Ferrimagnets\nKouki Nakata1,\u0003and Kei Suzuki1,y\n1Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, Ibaraki 319-1195, Japan\n(Dated: March 2, 2023)\nQuantum fluctuations are the key concepts of quantum mechanics. Quantum fluctuations of\nquantum fields induce a zero-point energy shift under spatial boundary conditions. This quantum\nphenomenon, called the Casimir effect, has been attracting much attention beyond the hierarchy\nof energy scales, ranging from elementary particle physics to condensed matter physics together\nwith photonics. However, the application of the Casimir effect to spintronics has not yet been in-\nvestigated enough, particularly to ferrimagnetic thin films, although yttrium iron garnet (YIG) is\none of the best platforms for spintronics. Here we fill this gap. Using the lattice field theory, we\ninvestigate the Casimir effect induced by quantum fields for magnons in insulating magnets and find\nthat the magnonic Casimir effect can arise not only in antiferromagnets but also in ferrimagnets\nincluding YIG thin films. Our result suggests that YIG, the key ingredient of magnon-based spin-\ntronics, can serve also as a promising platform for manipulating and utilizing Casimir effects, called\nCasimir engineering. Microfabrication technology can control the thickness of thin films and realize\nthe manipulation of the magnonic Casimir effect. Thus, we pave the way for magnonic Casimir\nengineering.\nIntroduction. —Toward efficient transmission of infor-\nmation that goes beyond what is offered by conven-\ntionalelectronics, thelasttwodecadeshaveseenasignifi-\ncant development of magnon-based spintronics [1], called\nmagnonics. The main aim of this research field is to use\nthe quantized spin waves, magnons, as a carrier of in-\nformation in units of the reduced Planck constant ~. A\npromising strategy for this holy grail is to explore insu-\nlating magnets. Thanks to the complete absence of any\nconducting metallic elements, insulating magnets are free\nfrom drawbacks of conventional electronics, such as sub-\nstantial energy loss due to Joule heating. This is the\nadvantage of insulating magnets. Thus, exploring quan-\ntum functionalities of magnons in insulating magnets is\na central task in the field of magnonics.\nQuantum fluctuations of photon fields induce a zero-\npoint energy shift, called the Casimir energy, under spa-\ntial boundary conditions. This Casimir effect is a funda-\nmentalphenomenonofquantumphysics, andtheoriginal\nplatform for the Casimir effect was the photon field [2–\n4], which is described by quantum electrodynamics. The\nconcept can be extended to various fields such as scalar,\nvector, tensor, and spinor fields. Nowadays, the Casimir\neffect has been attracting much attention beyond the hi-\nerarchyofenergyscales, rangingfromelementaryparticle\nphysics to condensed matter physics [5, 6]. As an exam-\nple, see Refs. [7–14] for Casimir effects in magnets [15].\nHowever, the application of the Casimir effect to spin-\ntronics has not yet been studied enough, particularly to\nferrimagnetic thin films (see Fig. 1), although yttrium\niron garnet (YIG) [16] has been playing a central role in\nspintronics.\nHere we fill this gap. In terms of the lattice field the-\nory, we investigate the Casimir effect induced by quan-\n\u0003nakata.koki@jaea.go.jp; equal contribution.\nyk.suzuki.2010@th.phys.titech.ac.jp; equal contribution.\nFIG. 1. Schematic of the ferrimagnetic thin film for the\nmagnonic Casimir effect. Two kinds of magnons (circles) arise\nfrom the alternating structure of up and down spins (arrowed\nlines). Wavy lines represent quantum-mechanical behaviors\nof magnons in the discrete energy.\ntum fields for magnons in insulating magnets and refer to\nit as the magnonic Casimir effect. We study the behav-\nior of the magnonic Casimir effect with a particular fo-\ncus on the thickness dependence of thin films, which can\nbeexperimentallycontrolledbymicrofabricationtechnol-\nogy [17, 18]. Then, we show that the magnonic Casimir\neffect can arise not only in antiferromagnets (AFMs) but\nalso in ferrimagnets with realistic model parameters for\nYIG thin films. Our study indicates that YIG, an ideal\nplatform for magnonics, can serve also as a key ingredi-\nent of Casimir engineering [19] which aims at exploring\nquantum-mechanical functionalities of nanoscale devices.\nAntiferromagnets. —We consider insulating AFMs de-\nscribed by the quantum Heisenberg Hamiltonian which\nhas U (1)symmetry about the quantization axis andarXiv:2205.13802v3 [quant-ph] 1 Mar 20232\nstudy the behavior of the magnonic Casimir effect with\na focus on the thickness dependence. The AFM is a\ntwo-sublattice system, and the ground state has the Néel\nmagnetic order [20]. From the spin-wave theory with the\nBogoliubov transformation, elementary excitations are\ntwo kinds of magnons designated by the index \u001b=\u0006\nhaving the spin angular momentum \u001b~. Owing to the\nU(1)symmetry, the Hamiltonian can be recast into the\ndiagonal form with the magnon energy dispersion for the\nwave number k= (kx;ky;kz)as\u000f\u001b;kwhere the total spin\nangular momentum of magnons is conserved. Two kinds\nof magnons ( \u001b=\u0006) are in degenerate states. Hence, we\nstudy the low-energy magnon dynamics of the insulat-\ning AFM by using the quantum field theory of complex\nscalar fields, i.e., the complex Klein-Gordon field the-\nory [21, 22]. Then, we can see that there exists a zero-\npoint energy [23]. This is the origin of the Casimir effect.\nNote that throughout this study, we focus on clean insu-\nlating magnets and work under the assumption that the\ntotal spin along the quantization direction is conserved\nand thus remains a good quantum number.\nThrough the lattice regularization, the Casimir energy\nECasis defined as the difference between the zero-point\nenergyEint\n0for the continuous energy \u000f\u001b;kand the one\nEsum\n0for the discrete energy \u000f\u001b;k;nwithn2Z. In two-\nsublattice systems, such as AFMs and ferrimagnets (see\nFig. 1), the wave numbers on the lattice are replaced by\n(kja)2!2[1\u0000cos(kja)]in thejdirection for j=x;y;z,\nwhereais the length of a magnetic unit cell. Here, by\ntaking the Brillouin zone (BZ) into account, we set the\nboundary condition for the zaxis in wave number space\nk= (kx;ky;kz)askz!\u0019n=Lz, i.e.,kza!\u0019n=Nz,\nwhereLz:=aNzis the thickness of magnets, Nj2N\nis the number of magnetic unit cells in the jdirection\nforj=x;y;z, andn= 1;2;:::;2NforN2N. The\nnumberofunitcellsonthe xyplaneis 4NxNy,andthatof\nmagneticunitcellsis NxNy. Thus, themagnonicCasimir\nenergy per the number of magnetic unit cells NxNyon\nthe surface for Nz=Nis described as [24–28]\nECas:=Esum\n0(N)\u0000Eint\n0(N); (1a)\nEsum\n0(N) =X\n\u001b=\u0006Z\nBZd2(k?a)\n(2\u0019)2\"\n1\n2\u00101\n22NX\nn=1j\u000f\u001b;k;nj\u0011#\n;(1b)\nEint\n0(N) =X\n\u001b=\u0006Z\nBZd2(k?a)\n(2\u0019)2\"\n1\n2NZ\nBZd(kza)\n2\u0019j\u000f\u001b;kj#\n;\n(1c)\nwherek?:=q\nk2x+k2y,d2(k?a) =d(kxa)d(kya), the\nintegral is over the first BZ, and the factor 1=2arises\nfrom the zero-point energy for the scalar field. Since\nthe constant terms which are independent of the wave\nnumber do not contribute to the Casimir energy, we drop\nthemthroughoutthisstudy. Toseethedependenceofthe\nCasimir energy ECason the thickness of magnets Lz:=\naNz, it is convenient to introduce the rescaled Casimir\n−1−0.8−0.6−0.4−0.2 0\n 0 2 4 6 8 10 12 14 16 18 20Antiferromagnetic thin film\nNzCasimir energy ECas [meV]\nNz: Thickness of magnet in units of aδ=2.0×10−3\nδ=1.0×10−3\nδ=0 −1−0.8−0.6−0.4−0.2 0\n 0 5 10 15 20CCas[3]=Nz3ECas [meV]FIG. 2. Plots of the magnonic Casimir energy ECasand its\ncoefficient (inset) C[3]\nCas=ECas\u0002N3\nzin the AFM [see Eq. (3a)]\nas a function of Nzfor the thickness of magnets Lz=aNz.\nenergyC[b]\nCasin terms of Nb\nzforb2Ras\nC[b]\nCas:=ECas\u0002Nb\nz: (2)\nThen, we refer to C[b]\nCasas the magnonic Casimir coeffi-\ncient in the sense that ECas=C[b]\nCasN\u0000b\nz.\nHere, we consider magnons in AFMs on a cubic lattice\nwith the energy dispersion \u000f\u001b;k=\u000fAFM\n\u001b;k[29]:\n\u000fAFM\n\u001b;k=~!0r\n\u000e+\u0010ka\n2\u00112\n; (3a)\n~!0:=2p\n3JS; (3b)\n\u000e:=3h\u0010K\n6J\u00112\n+ 2\u0010K\n6J\u0011i\n; (3c)\nwherek:=jkj,J >0parametrizes the exchange interac-\ntion between the nearest-neighbor spins of the spin quan-\ntum number S, andK\u00150is the easy-axis anisotropy,\nwhich provides the magnon energy gap and ensures the\nNéel magnetic order. Two kinds of magnons ( \u001b=\u0006)\nare in degenerate states. In the absence of the spin\nanisotropy, K= 0, the energy gap vanishes \u000e= 0, and\nthegaplessmagnonmodebehaveslikearelativisticparti-\nclewiththelinearenergydispersion. Fromtheresultsob-\ntainedinRefs.[30–32],weroughlyestimatethemodelpa-\nrameter values for Cr 2O3, as an example, as follows [29]:\nJ= 15meV,S= 3=2,K= 0:03meV, anda= 0:496 07\nnm. These parameter values provide ~!0\u001877:94meV\nand\u000e\u00182\u000210\u00003.\nFigure 2 shows that the magnonic Casimir effect arises\nin the thin film of the AFM. The magnonic Casimir en-\nergyECasof the magnitude O(10\u00002)meV is induced\nforNz\u00152. Even in the presence of the magnon en-\nergy gap\u000e6= 0, the absolute value amounts to O(10\u00002)\nmeV and decreases as the magnon energy gap increases.\nThus, the magnonic Casimir energy takes a maximum3\nabsolute value in the gapless mode \u000e= 0, where the\nmagnon behaves like a relativistic particle with the lin-\near energy dispersion. We remark that in the case of\nthe gapped magnon modes, the absolute value of the\nmagnonicCasimircoefficient C[3]\nCas=ECas\u0002N3\nzdecreases\nas the thickness of the thin film increases. This behav-\nior is similar to the Casimir effect known for massive\ndegrees of freedom [33, 34]. In the case of the gapless\nmode, the magnonic Casimir coefficient C[3]\nCasapproaches\nasymptotically to a constant value, and the magnonic\nCasimir energy exhibits the behavior of ECas/1=N3\nzas\nthe thickness increases. The asymptotic value of C[3]\nCas\nfor the gapless magnon mode \u000e= 0given in the nu-\nmerical result (see Fig. 2) is estimated approximately as\n(\u0000\u00192=720)\u0002(~!0=2)\u0018\u00000:5341meV from an analyt-\nical calculation. The factor of \u0000\u00192=720is well known\nas the analytic solution for the conventional Casimir\neffect of a massless complex scalar field in continuous\nspace [34]. Thus, although the magnonic Casimir effect\nis realized on the lattice, it is qualitatively and quanti-\ntatively analogous to the continuous counterpart, except\nfora-dependent lattice effects.\nFerrimagnets. —We develop the study of AFMs into\nferrimagnets where the ground state has an alternating\nstructure of up and down spins on a cubic lattice (see\nFig. 1). In contrast to AFMs, the spin quantum num-\nber on the two-sublattice is different from each other in\nferrimagnets. Hence, the degeneracy for two kinds of\nmagnons (\u001b=\u0006) is intrinsically lifted. In ferrimagnetic\nthinfilms, dipolarinteractionsduetothenonzeromagne-\ntization play a key role. Still, at low temperatures where\nthe magnon-magnon interaction of the fourth order in\nmagnon operators is negligibly small [35], the number of\nmagnons and the total spin angular momentum are con-\nserved, and the Hamiltonian for the ferrimagnetic thin\nfilm can be diagonalized with the magnon energy disper-\nsion\u000f\u001b;k=\u000fferri\n\u001b;k.\nHere, we consider magnons in the thin film of clean\ninsulating ferrimagnets on a cubic lattice subjected to\nin-plane magnetic fields at such low temperatures. Still,\ndue to the competition between dipolar and exchange\ninteractions, the minimum energy point shifts from the\nzero mode of magnons, k= 0, to a finite wave number\nmode which is characterized by the thickness of the thin\nfilmLz=aNz(see Fig. 1).\nThe magnon energy dispersion along the in-plane di-\nrection is provided in Refs. [36, 37], whereas the disper-\nsion along the zaxis in the thin film has not yet been\nestablished [38]. Hence, taking into account the compe-\ntition between dipolar and exchange interactions in the\nthin film, we phenomenologically assume the behavior\nthat the power of kz,l2R, approaches asymptotically\ntol= 2in the bulk limit, whereas it slightly differs from\nl= 2as long as we consider the thin film (see Fig. 1).\nUsing this assumption and Refs. [36, 37], the magnon\n−120−100−80−60−40−20 0 20\n 0 2 4 6 8 10 12 14 16 18 20Ferrimagnetic thin film\nNzCasimir energy ECas [µeV]\nNz: Thickness of magnet in units of al=2.1\nl=2.0l=1.99\nl=1.9\n−100−50 0 50 100\n 0 5 10 15 20CCas[l]=NzlECas [µeV]FIG. 3. Plots of the magnonic Casimir energy ECasand its\ncoefficient (inset) C[l]\nCas=ECas\u0002Nl\nzforl= 2:1,l= 2:0,l=\n1:99, and l= 1:9in the ferrimagnetic thin film [see Eq. (4a)]\nas a function of Nzfor the thickness of magnets Lz=aNz\nwith the model parameter values for YIG of Dz=D.\nenergy dispersions are\n\u000fferri\n\u001b;k=r\n\u001bH0+ \u0001\u001b+D\na2(k?a)2+Dz\na2(jkzja)l(4a)\n\u0002r\n\u001bH0+ \u0001\u001b+D\na2(k?a)2+Dz\na2(jkzja)l+\u001b~!MFk;\nFk:=Pk?(1\u0000Pk?)\u001b~!M\n\u001bH0+ \u0001\u001b+D\na2(k?a)2+Dz\na2(jkzja)l\u0010kx\nk?\u00112\n+1\u0000Pk?\u0010ky\nk?\u00112\n; (4b)\nPk?:=1\u00001\u0000e\u0000k?Lz\nk?Lz; (4c)\nwhere the external magnetic field is applied along the y\naxis (see Fig. 1) and \u001bH0represents the resulting Zee-\nman energy, \u0001\u001b\u00150is the (intrinsic) magnon energy\ngap in ferrimagnets, D(z)>0is the spin stiffness con-\nstant,k?:=q\nk2x+k2y,~!M:= 4\u0019\rMswith the sat-\nurated magnetization density Msand the gyromagnetic\nratio\r, and the termFkis responsible for the shift of\nthe minimum energy point from the zero mode to a finite\nwave number mode due to the competition between dipo-\nlar and exchange interactions in the ferrimagnetic thin\nfilm: The first term of Fk[see Eq. (4b)] reproduces the\nDamon-Eshbach magnetostatic surface mode [39], and\nthe last term reproduces the backward volume magneto-\nstatic mode [39].\nFrom the results obtained in Refs. [40–42], the model\nparameter values for YIG thin films are estimated as fol-\nlows:D=a2\u00183:376 45meV [35] with a= 1:2376nm [43],\nH0\u00188:103 73\u0016eV [35], and ~!M\u001820:3369\u0016eV [35].\nThen, we estimate the magnon energy gap \u0001\u001bas [44]\n\u0001\u001b=\u0000\u0000\u0001\u001b=+\u001839:848 81meV with \u0001\u001b=+\u00182:131 91\nmeV and \u0001\u001b=\u0000\u001841:980 72meV, which satisfy the con-4\ndition \u0001\u001b\u001d~!M. In this condition, we study the low-\nenergy magnon dynamics of the ferrimagnetic thin film\nby using the quantum field theory of real scalar fields,\ni.e., the real Klein-Gordon field theory [21, 22]. Then,\nwe can see that there exists a zero-point energy [35].\nThe magnonic Casimir energy through the lattice reg-\nularization is given as Eq. (1a). We remark that the\nzero-point energy arises from quantum fluctuations and\ndoes exist even at zero temperature. The zero-point en-\nergy defined at zero temperature does not depend on the\nBose-distribution function [Eqs. (1b) and (1c)]. Hence,\nnot only the low-energy mode ( \u001b= +) but also the high-\nenergy mode ( \u001b=\u0000) contribute [45] to the magnonic\nCasimir energy.\nUnderthephenomenologicalassumptionthatthevalue\nofl[see Eq. (4a)] approaches asymptotically to l= 2in\nthe bulk limit, in this work focusing on the thin film,\nwe study the behavior of the magnonic Casimir effect\nby changing the value slightly from l= 2. As exam-\nples, we consider the cases of l= 2:1,l= 1:99, and\nl= 1:9. Figure 3 shows that the magnonic Casimir ef-\nfect arises in the ferrimagnetic thin film. There is the\nmagnonic Casimir energy ECasof the magnitude O(1),\nO(10), andO(10)\u0016eV forl= 1:99,l= 1:9, andl= 2:1,\nrespectively, in Nz\u00152. As the thickness increases, the\nmagnonic Casimir coefficient C[l]\nCasapproaches asymptot-\nically to a constant value, and the magnonic Casimir en-\nergy exhibits the behavior of ECas/1=Nl\nz. We also\nfind from Fig. 3 that the sign of the magnonic Casimir\ncoefficient and energy for l= 2:1is positive C[2:1]\nCas =\nECas\u0002N2:1\nz>0inNz\u00152, whereas that for l= 1:9is\nnegativeC[1:9]\nCas=ECas\u0002N1:9\nz<0. This means that the\nCasimir force works in the opposite direction.\nNote that even if l= 2, the magnonic Casimir effect\narises in the ferrimagnetic thin film. We have numer-\nically confirmed that although the value is small, there\ndoes exist the magnonic Casimir energy of the magnitude\njECasj\u0014O(0:1)neV forl= 2. This strong suppression of\nthe Casimir energy is a general property of the Casimir\neffect for quadratic dispersions on the lattice [28], and\nthe survival values originate from the dipolar interaction\nin the ferrimagnetic thin film.\nWe remark that as long as we consider thin films,\nthe value of Dzcan differ from D. Even in that case,\nthe magnonic Casimir effect arises. When the value\nofDzchanges from Dto0:8Das an example, the\nmagnonic Casimir energy ECasincreases approximately\nby0:8times. For more details about its dependence on\nthe parameters Dzandl, see the Supplemental Material.\nProposal for experimental observation. —Themagnonic\nCasimir energy of the ferrimagnetic thin film depends\nstronglyonexternalmagneticfieldsthroughZeemancou-\npling as in Eq. (4a) and contributes to magnetization of\nmagnets, whereas the photon and phonon Casimir ef-\nfects [46] do not usually. On the other hand, in the\npresenceofmagnetostriction[47–50], thephononCasimir\neffect is influenced by magnetostriction, and its correc-\ntion for the phonon Casimir energy depends on magneticfields and contributes to magnetization. However, such a\ncontribution to magnetization from the phonon Casimir\neffect should be negligibly small by the factor of 10\u00006\ncompared with that from the magnonic Casimir effect\nof ferrimagnets because the magnetostriction constant\n(i.e., the correction for the lattice constant) is known to\nbe10\u00006for YIG [47, 48]. Hence, even in the presence\nof magnetostriction, the magnonic Casimir effect can be\ndistinguished from the others. Thus, we expect that our\ntheoreticalprediction, themagnonicCasimireffectinfer-\nrimagnets, can be experimentally observed through mea-\nsurement of magnetization and its film thickness depen-\ndence. For more details, see the Supplemental Material.\nForobservation, afewcommentsareinorder. First, we\nremark on edge/surface magnon modes. The magnonic\nCasimir effect in our setup (see the thin film of Fig. 1)\nis induced by magnon fields with wave numbers kzdis-\ncretized by small Nz, and its necessary condition is a\nkz-dependent dispersion relation under the discretization\nofkz. Throughout this study, we consider thin films of\nNz\u001cNx;Ny. Even if additional edge/surface magnon\nmodesexistaswellastheDamon-Eshbachmagnetostatic\nsurface mode and the backward volume magnetostatic\nmode [see Eq. (4b)], they are confined only on the x-\nyplane. Then, their wave number in the zdirection\nis always zero, i.e., kz= 0, and its energy dispersion\nrelation is independent of kz. Since a kz-independent\ndispersion relation cannot induce the Casimir effect, the\nedge/surface modes cannot contribute to the magnonic\nCasimir effect. In this sense, our magnonic Casimir effect\nis not affected by the existence of edge/surface magnon\nmodes.\nNote that details of the edge condition, such as the\npresence or absence of disorder, may change the bound-\nary condition for the wave function of magnons, but the\nexistence of the magnonic Casimir effect remains un-\nchanged. Even if there is a change in the spectrum near\nthe edge, the magnonic Casimir effect is little influenced\nas long as one does not assume an ultrathin film such\nasNz= 1;2;3. In this sense, we expect that the fol-\nlowing size of thin films is appropriate for observation of\nour prediction: Nz\u001810, i.e., the film thickness of YIG\nisLz=aNz\u001812:376nm. Note that microfabrication\ntechnology [17, 18] can control the thickness of thin films\nand realize the manipulation of the magnonic Casimir\neffect.\nNext, we remark on the magnon band structure.\nSince our magnonic Casimir effect is induced by the kz-\ndependent dispersion, its Casimir energy of ferrimagnets\nis mainly characterized by the Dz-term in Eq. (4a), i.e.,\nDz(jkzja)l. Hence, we have investigated its dependence\non bothlandDz(see Fig. 3 and the Supplemental Mate-\nrial). Even if the magnon band structure is affected due\nto some reasons, the magnonic Casimir effect of ferrimag-\nnets is little influenced by other details of the magnon\nband structure except for landDz.\nLastly, we remark on thermal effects. At nonzero tem-\nperature, thermal contributions to the Helmholtz free en-5\nergy arise. However, at low temperatures compared to\n\u000f\u001b;k[51], the thermal contribution is exponentially sup-\npressed due to the Boltzmann factor and becomes negli-\ngibly small. Hence, at such low temperatures, the contri-\nbution of the magnonic Casimir energy given as Eq. (1a)\nis dominant.\nConclusion. —We have shown that the magnonic\nCasimir effect can arise not only in antiferromagnets but\nalso in ferrimagnets with realistic model parameters for\nYIG. Since the lifetime of magnons in YIG thin films is\nthelongestamongknownmaterials, andmagnonsexhibit\nlong-distance transport over centimeter distances [52],\nYIG is the key ingredient of magnonics [1, 16], which\nhas already realized the magnon transistor [53]. Ourresult suggests that YIG can serve also as a promis-\ning platform for Casimir engineering [19]: Because the\nmagnonic Casimir effect contributes to the internal pres-\nsure of thin films, it will provide the new principles of\nnanoscale devices such as highly sensitive pressure sen-\nsors and magnon transistors. Thus, our study paves the\nway for magnonic Casimir engineering.\nWe thank Yasufumi Araki, Yoshinori Haga, Masaki\nImai, Se Kwon Kim, and Katsumasa Nakayama for fruit-\nful discussions. We acknowledge support by Leading Ini-\ntiative for Excellent Young Researchers, MEXT, Japan\n(K.N.), by JSPS KAKENHI Grants No. JP20K14420\n(K. N.), No. JP22K03519 (K. N.), No. JP17K14277 (K.\nS.), and No. JP20K14476 (K. S.).\n[1] A. V. Chumak, V. I. Vasyuchka, A. A. Serga, and\nB. Hillebrands, Nat. Phys. 11, 453 (2015).\n[2] H. B. G. Casimir, Proc. Kon. Ned. Akad. Wetensch. 51,\n793 (1948).\n[3] S. K. Lamoreaux, Phys. Rev. Lett. 78, 5 (1997).\n[4] S. K. Lamoreaux, Phys. Rev. Lett. 81, 5475(E) (1998).\n[5] G. L. Klimchitskaya, U. Mohideen, and V. M.\nMostepanenko, Rev. Mod. Phys. 81, 1827 (2009),\narXiv:0902.4022.\n[6] K. A. Milton, J. Phys. A 37, R209 (2004), arXiv:hep-\nth/0406024.\n[7] H. Neuberger and T. Ziman, Phys. Rev. B 39, 2608\n(1989).\n[8] P. Hasenfratz and F. Niedermayer, Z. Phys. B 92, 91\n(1993), arXiv:hep-lat/9212022.\n[9] L. P. Pryadko, S. Kivelson, and D. W. Hone, Phys. Rev.\nLett.80, 5651 (1998), arXiv:cond-mat/9711129.\n[10] Z. Z. Du, H. M. Liu, Y. L. Xie, Q. H. Wang, and J.-M.\nLiu, Phys. Rev. B 92, 214409 (2015), arXiv:1506.05211.\n[11] R. Cheng, D. Xiao, and J.-G. Zhu, Phys. Rev. Lett. 121,\n207202 (2018), arXiv:1802.07867.\n[12] E. B. Kolomeisky, H. Zaidi, L. Langsjoen, and J. P. Stra-\nley, Phys. Rev. A 87, 042519 (2013), arXiv:1110.0421.\n[13] A. Roldán-Molina, M. J. Santander, A. S. Nunez, and\nJ. Fernández-Rossier, Phys. Rev. B 92, 245436 (2015),\narXiv:1502.01950.\n[14] B. A. Ivanov, D. D. Sheka, V. V. Kryvonos, and F. G.\nMertens, Phys. Rev. B 75, 132401 (2007).\n[15] See also Ref. [54] for an analog of the dynamical Casimir\neffect [55–57] with magnon excitations in a spinor Bose-\nEinstein condensate.\n[16] A. Serga, A. Chumak, and B. Hillebrands, J. Phys. D\n43, 264002 (2010).\n[17] G. Schmidt, C. Hauser, P. Trempler, M. Paleschke, and\nE. T. Papaioannou, Phys. Stat. Sol. B 257, 1900644\n(2020), arXiv:2004.09094.\n[18] Y. Yang, T. Liu, L. Bi, and L. Deng, J. Alloys Compd.\n860, 158235 (2021).\n[19] T.Gong, M.R.Corrado, A.R.Mahbub, C.Shelden, and\nJ. N. Munday, Nanophotonics 10, 523 (2021).\n[20] As an example, Ref. [58] studied magnon dynamics at\nthe zigzag edge of a honeycomb lattice with long-ranged\nNéel magnetic order.\n[21] M. E. Peskin and D. V. Schroeder, An Introduction ToQuantum Field Theory (Westview Press, Boulder, CO,\n1995).\n[22] Z. F. Ezawa, Quantum Hall Effects: Recent Theoretical\nand Experimental Developments, 3rd ed.(World Scien-\ntific Publishing Co. Pte. Ltd., Singapore, 2013).\n[23] P. W. Anderson, Phys. Rev. 86, 694 (1952).\n[24] A. Actor, I. Bender, and J. Reingruber, Fortschr. Phys.\n48, 303 (2000), arXiv:quant-ph/9908058.\n[25] M. Pawellek, arXiv:1303.4708 .\n[26] T. Ishikawa, K. Nakayama, and K. Suzuki, Phys. Lett.\nB809, 135713 (2020), arXiv:2005.10758.\n[27] T. Ishikawa, K. Nakayama, and K. Suzuki, Phys. Rev.\nRes.3, 023201 (2021), arXiv:2012.11398.\n[28] K. Nakayama and K. Suzuki, arXiv:2204.12032 .\n[29] K. Nakata, S. K. Kim, J. Klinovaja, and D. Loss, Phys.\nRev. B 96, 224414 (2017), arXiv:1707.07427.\n[30] J. O. Artman, J. C. Murphy, and S. Foner, Phys. Rev.\n138, A912 (1965).\n[31] Y. Kota and H. Imamura, Appl. Phys. Express 10,\n013002 (2017).\n[32] R. E. Newnham and Y. M. De Haan, Z. Kristallogr.\nCryst. Mater. 117, 235 (1962).\n[33] P. Hays, Ann. Phys. (N. Y.) 121, 32 (1979).\n[34] J. Ambjørn and S. Wolfram, Ann. Phys. (N. Y.) 147, 1\n(1983).\n[35] I. S. Tupitsyn, P. C. E. Stamp, and A. L. Burin, Phys.\nRev. Lett. 100, 257202 (2008).\n[36] B. A. Kalinikos and A. N. Slavin, J. Phys. C 19, 7013\n(1986).\n[37] K. Y. Guslienko, R. W. Chantrell, and A. N. Slavin,\nPhys. Rev. B 68, 024422 (2003).\n[38] S. M. Rezende, Fundamentals of Magnonics (Springer,\nHeidelberg, 2020).\n[39] R. W. Damon and J. R. Eshbach, J. Phys. Chem. Solids\n19, 308 (1961).\n[40] R. Pauthenet, Ann. Phys. (Paris) 13, 424 (1958).\n[41] M. A. Gilleo and S. Geller, Phys. Rev. 110, 73 (1958).\n[42] M. Sparks, Ferromagnetic-Relaxation Theory (McGraw-\nHill, New York, 1964).\n[43] S. Geller and M. A. Gilleo, J. Phys. Chem. Solids 3, 30\n(1957).\n[44] The magnon energy dispersions and their temperature\ndependence in YIG were measured by inelastic neutron\nscattering [59–65]. We estimate the magnon energy gap6\n\u0001\u001bby applying the model [66] of the effective block spins\nto YIG [35, 63]. The theoretical estimate for the value,\n\u0001\u001b=\u0000\u0000\u0001\u001b=+\u001839:848 81meV, is consistent with the\nexperimental data [62].\n[45] Higher energy bands than those of Eq. (4a) also con-\ntribute to the magnonic Casimir energy. However, the\ncontribution becomes smaller as the shape of the bands\nis flatter. Numerical calculations of Refs. [63–65] show\nthat higher energy bands tend to be flat. Thus, we ex-\npect that the magnonic Casimir energy is dominated by\nthe two bands of Eq. (4a).\n[46] See Refs. [67–71] for the Casimir effect of phonons and,\ne.g., Ref [72] for the dynamical one.\n[47] A. Smith and R. Jones, J. Appl. Phys. 34, 1283 (1963).\n[48] E. R. Callen, A. E. Clark, B. DeSavage, W. Coleman,\nand H. B. Callen, Phys. Rev. 130, 1735 (1963).\n[49] K. Dudko, V. Eremenko, and L. Semenenko, Phys. Stat.\nSol.43, 471 (1971).\n[50] R. Yacovitch and Y. Shapira, Physica (Amsterdam)\n86B+C, 1126 (1977).\n[51] Note that even at such low temperatures, magnons do\nnot form Bose-Einstein condensates in equilibrium [73].\n[52] L. J. Cornelissen, J. Liu, R. A. Duine, J. B. Youssef,\nand B. J. Van Wees, Nat. Phys. 11, 1022 (2015),\narXiv:1505.06325.\n[53] A. V. Chumak, A. A. Serga, and B. Hillebrands, Nat.\nCommun. 5, 4700 (2014).\n[54] H. Saito and H. Hyuga, Phys. Rev. A 78, 033605 (2008),\narXiv:0805.2210.\n[55] G. T. Moore, J. Math. Phys. (N.Y.) 11, 2679 (1970).\n[56] V. Dodonov, Phys. Scr. 82, 038105 (2010),\narXiv:1004.3301.\n[57] P. D. Nation, J. R. Johansson, M. P. Blencowe, and\nF. Nori, Rev. Mod. Phys. 84, 1 (2012), arXiv:1103.0835.\n[58] W.-M. Huang, T. Hikihara, Y.-C. Lee, and H.-H. Lin,\nSci. Rep. 7, 43678 (2017), arXiv:1107.1102.\n[59] J. Plant, J. Phys. C 10, 4805 (1977).\n[60] J. Plant, J. Phys. C 16, 7037 (1983).\n[61] S.-i. Shamoto, T. U. Ito, H. Onishi, H. Yamauchi, Y. In-\namura, M. Matsuura, M. Akatsu, K. Kodama, A. Nakao,\nT. Moyoshi, et al., Phys. Rev. B 97, 054429 (2018),\narXiv:1705.02167.\n[62] Y. Nambu, J. Barker, Y. Okino, T. Kikkawa, Y. Shiomi,\nM. Enderle, T. Weber, B. Winn, M. Graves-Brook, J. M.\nTranquada, et al., Phys. Rev. Lett. 125, 027201 (2020),\narXiv:1911.11968.\n[63] Y. Nambu and S.-i. Shamoto, J. Phys. Soc. Jpn. 90,\n081002 (2021), arXiv:2106.15752.\n[64] A. J. Princep, R. A. Ewings, S. Ward, S. Tóth, C. Dubs,\nD. Prabhakaran, and A. T. Boothroyd, npj Quantum\nMater. 2, 63 (2017).\n[65] O. I. Gorbatov, G. Johansson, A. Jakobsson,\nS. Mankovsky, H. Ebert, I. Di Marco, J. Minár,\nand C. Etz, Phys. Rev. B 104, 174401 (2021).\n[66] K. Nakata and S. Takayoshi, Phys. Rev. B 102, 094417\n(2020), arXiv:2004.03353.\n[67] M. Schecter and A. Kamenev, Phys. Rev. Lett. 112,\n155301 (2014), arXiv:1307.4409.\n[68] A. I. Pavlov, J. van den Brink, and D. V. Efremov, Phys.\nRev. B 98, 161410 (2018), arXiv:1809.11071.\n[69] A. I. Pavlov, J. van den Brink, and D. V. Efremov, Phys.\nRev. B 100, 014205 (2019), arXiv:1812.09004.\n[70] A. Rodin, Phys. Rev. B 100, 195403 (2019),\narXiv:1908.02006.[71] G. Lee and A. Rodin, Phys. Rev. B 103, 195434 (2021),\narXiv:2012.11082.\n[72] X. Wang, W. Qin, A. Miranowicz, S. Savasta,\nand F. Nori, Phys. Rev. A 100, 063827 (2019),\narXiv:1902.09910.\n[73] Yu. M. Bunkov and G. E. Volovik, arXiv:1003.4889 .7\nSupplemental Material\nIn this Supplemental Material, first, we provide some\ndetails about the dependence of the magnonic Casimir\neffect on the parameters landDzin ferrimagnetic thin\nfilms. Next, we remark on its film thickness dependence.\nThen, we provide another point of view for its robust-\nness against disorder effects. Lastly, we comment on the\ndistinction between the Casimir effect and the thermal\nCasimir effect.\nI. THE PARAMETER l- AND Dz-DEPENDENCE\nIn the main text, we have studied the magnonic\nCasimirenergy ECasandthecoefficient C[l]\nCas=ECas\u0002Nl\nz\nforl= 2:1,l= 2:0,l= 1:99, andl= 1:9in the ferrimag-\nnetic thin film by using the model parameter values for\nYIG with fixed Dz=D. Here, we provide more details\nabout its dependence on the parameters landDz.\nFirst, we consider the cases of l= 1:5andl= 1:0with\nsettingDz=D. Figure S1 shows that the magnonic\nCasimir effect still arises in the ferrimagnetic thin film.\nThere is the magnonic Casimir energy ECasof the mag-\nnitudeO(10\u00002)meV,O(10\u00001)meV, andO(10\u00001)meV\nforl= 1:9,l= 1:5, andl= 1:0, respectively, in Nz\u00152.\nAs the value of ldecreases from l= 2and approaches\ntol= 1, the magnitude of the magnonic Casimir energy\nincreases. Note that it amounts to O(10\u00001)meV even in\nNz=O(10)forl= 1:0. As the thickness increases, the\nmagnonic Casimir coefficient C[l]\nCasapproaches asymptot-\nically to a constant value and the magnonic Casimir en-\nergy exhibits the behavior of ECas/1=Nl\nz.\nThen, we consider the cases of Dz=D= 0:3,Dz=D=\n0:5, andDz=D= 0:8by fixingl= 1:99. Figure S2 shows\nthat the magnonic Casimir effect still arises in the ferri-\nmagnetic thin film. When the value of Dzchanges from\nDto0:8Das an example, the magnonic Casimir energy\nECasincreases approximately by 0:8times. Thus, the\nvalue ofECasis approximately proportional to Dz.\nII. REMARKS ON THE THICKNESS\nDEPENDENCE OF MAGNETIZATION\nIn the main text, we have remarked that our predic-\ntion, the magnonic Casimir effect in ferrimagnets, can\nbe observed through measurement of magnetization and\nits film thickness dependence. Here, we add an expla-\nnation about it. At zero temperature, the Helmholtz\nfree energy of magnon fields in thin films (i.e., the\nsum over discrete kz) isEsum\n0(Nz)NxNy, and that un-\nder the bulk approximation (i.e., the integral with re-\nspect to continuous kz) isEint\n0(Nz)NxNy[see Eqs. (1a)-\n(1c)]. The difference between them is characterized by\nthe magnonic Casimir energy ECasasEsum\n0(Nz)NxNy=\nEint\n0(Nz)NxNy+ECasNxNy, where the magnon energy\n−1−0.8−0.6−0.4−0.2 0\n 0 2 4 6 8 10 12 14 16 18 20Ferrimagnet\nNzCasimir energy ECas [meV]\nNz: Thickness of magnet in units of al=2.0\nl=1.9l=1.5\nl=1.0\n−1−0.8−0.6−0.4−0.2 0\n 0 5 10 15 20CCas[l]=NzlECas [meV]FIG. S1. Plots of the magnonic Casimir energy ECasand\nthe coefficient (inset) C[l]\nCas=ECas\u0002Nl\nzforl= 2:0,l= 1:9,\nl= 1:5, and l= 1:0intheferrimagneticthinfilmasafunction\nofNzfor the thickness of magnets Lz=aNzunder the model\nparameter values for YIG with fixed Dz=D.\n−14−12−10−8−6−4−2 0\n 0 2 4 6 8 10 12 14 16 18 20Ferrimagnet\n(l=1.99)\nNzCasimir energy ECas [µeV]\nNz: Thickness of magnet in units of aDz=0.3D\nDz=0.5DDz=0.8D\nDz=D\n−14−12−10−8−6−4−2 0\n 0 5 10 15 20CCas[l]=NzlECas [µeV]\nFIG. S2. Plots of the magnonic Casimir energy ECasand the\ncoefficient (inset) C[l]\nCas=ECas\u0002Nl\nzforDz=D= 0:3,Dz=D=\n0:5,Dz=D= 0:8, and Dz=D= 1:0in the ferrimagnetic thin\nfilmasafunctionof Nzforthethicknessofmagnets Lz=aNz\nunder the model parameter values for YIG with l= 1:99.\ndispersion of ferrimagnets (i.e., magnets including dipo-\nlar interactions) is Eq. (4a). Note that the magnetic-\nfield derivative (i.e., H0-derivative) of the Helmholtz free\nenergy is magnetization. Then, magnetization of thin\nfilms consists of two parts: The bulk component and the\nmagnonic Casimir energy. Since Eint\n0(Nz)/Nz, whereas\nECas/1=(Nz)l, magnetization of thin films exhibits a\ndifferentNz-dependence from the bulk component, and\nits difference is characterized by the magnonic Casimir\nenergy. In other words, magnetization of thin films ex-\nhibits a different film thickness dependence from the bulk\ncomponent due to the magnonic Casimir effect. Hence,\nour prediction, the magnonic Casimir effect in ferrimag-8\nnetic thin films (i.e., magnetic thin films including dipo-\nlar interactions), can be observed through measurement\nof magnetization and its film thickness dependence.\nNote that if dipolar interactions are relevant also in\nantiferromagnets, its low-energy magnon dynamics is es-\nsentially described by Eq. (4a) given for ferrimagnets.\nThe only difference is that the spin quantum number for\neachsublatticeisidenticalinantiferromagnets, wherethe\n(intrinsic) magnon energy gap for each mode \u001b=\u0006can\nbe identical \u0001\u001b=+= \u0001\u001b=\u0000[see Eq. (4a)]. In this sense,\nitsCasimireffectexhibitsqualitativelythesamebehavior\nas Fig. 3.\nIII. REMARKS ON DISORDER EFFECTS\nIn the main text, we have remarked that details of the\nedge condition, such as the presence or absence of dis-\norder, may change the boundary condition for the wave\nfunction of magnons, but the existence of the magnonic\nCasimir effect remains unchanged. Here, we add a com-\nment on disorder effects. Since the magnonic Casimir\nenergy does not depend on the Bose-distribution func-\ntion [see Eqs. (1a)-(1c)], not only the low-energy magnon\nmode (\u001b= +) but also its high-energy mode ( \u001b=\u0000) in\nferrimagnets contributes to the magnonic Casimir effect.\nTherefore, it can be expected that as long as disorder\neffects on the bulk are weak enough that the high-energy\nmode is little influenced, the existence of the magnonic\nCasimir effect in ferrimagnets remains unchanged.\nIV. THERMAL CASIMIR EFFECT\nIn the main text, we have explained that thermal con-\ntributions to the Helmholtz free energy arise at nonzero\ntemperature. Here, we add a remark on it. Although its\nthermal contribution is called the “thermal Casimir en-\nergy”, there is a crucial distinction between the Casimir\neffect and the thermal Casimir effect: The zero-point en-\nergy, which is the key concept of quantum mechanics and\nplays a crucial role in the Casimir effect, is absent in the\nthermal Casimir effect. It should be noted that the zero-\npoint energy is one of the most striking phenomenon of\nquantum mechanics in the sense that there are no clas-\nsical analogs. The Casimir effect arises from the zero-\npoint energy due to quantum fluctuations and is not af-\nfected by temperatures, whereas the thermal Casimir ef-\nfect arises from thermal fluctuations and is exponentially\nsuppressed at low temperatures. The thermal Casimir\neffect vanishes at zero temperature, whereas the Casimir\neffect does exist even at zero temperature. Thus, there is\na significant distinction between the Casimir effect and\nthe thermal Casimir effect." }, { "title": "1510.05087v1.Ferrimagnetic_nanostructures_for_magnetic_memory_bits.pdf", "content": "1 \n Ferrimagnetic nanostructures for magnetic memory bits \nAUTHOR NAMES \nA. A. Ünal ,1 S. Valencia,1 D. Marchenko,1 K. J. Merazzo,2 F. Radu,1 M. Vázquez,2 and J. \nSánchez -Barriga1,* \nAUTHOR ADDRESS \n1Helmholtz -Zentrum Berlin für Materiali en und Energie, Albert -Einstein -Str. 15, 12489 \nBerlin, Germany \n2Instituto de Ciencia de Materiales de Madrid, CSIC, 28049 Madrid, Spain \nKEYWORDS Heat -assisted magnetic recording, bit- patterned media, magnetic data storage, \nferrimagnetic antidots \n \n \n 2 \n ABSTRACT \nIncreasing the magnetic data recording density requires reducing the size of the individual \nmemory elements of a recording layer as well as employing magn etic materials with \ntemperature- dependent functionalities. Therefore, it is predicted that t he near future of \nmagnetic data storage technology involves a combination of energy -assisted recording on \nnanometer -scale magnetic media. We present the potential of heat -assisted magnetic \nrecording on a patterned sample; a ferrimagnetic alloy composed of a rare earth and a transition metal, DyCo\n5, which is grown on a hexagonal -ordered nanohole array membrane. \nThe magnetization of the antidot array sample is out- of-plane oriented at room temperature \nand rotates towards in- plane upon heating above its spin- reorientation temp erature (TR) of \n~350 K, just above room temperature. Upon cooling back to room temperature (below T R), we \nobserve a well -defined and unexpected in- plane magnetic domain configuration modulating \nwith ~45 nm. We discuss the underlying mechanisms giving rise to this behavior by \ncomparing the magnetic properties of the patterned sample with the ones of its extended thin \nfilm counterpart . Our results pave the way for novel applications of ferrimagnetic antidot \narrays of superior functionality in magnetic nano -devices near room temperature. \nINTRODUCTION \nThe research on artificially engineered spintronic materials in the for m of layered thin films \nand low -dimensional structures has experienced a tremendous boost due to the versatility of \ntheir applications in computing technology. This is reflected by the development of magnetic \ntunneling based read- out technologies [ 1], magnetic sensors [ 2,3], and non- volatile \nmagnetoresistive random -access memory devices [ 4]. Because conventional magnetic data \nstorage relies on two -dimensional arrays of magnetic bits, developing higher density, cheaper \nand faster devices relies on reducing the lateral size of individual memory elements or data 3 \n storage bits [ 5]. Ultimately controlling the magnetic anisotropy and shape of nanostructures, \nas well as understanding their magnetic properties in reduced dimensions are key aspects that \nhave received a sustained interest over the past decades [ 6]. For instance, in antiferromagnetic \n(AF) materials, which have crucial impo rtance in exchange- bias systems, finite -size effects \nlead to a scaling of the magnetic properties such as the ordering temperature or the magnetic anisotropy [ 7], while in ferromagnetic (FM) materials control of the shape anisotropy can be \nachieved dependi ng on the geometry of the objects [ 8]. \nNanostructured magnetic materials often exhibit novel properties over their bulk \ncounterparts as a consequence of reduced atomic coordinates and modified density of states. Magnetic systems of low dimensionality like artificially grown or self -organized FM \nnanostructures [ 9,10], arrays of nanostructures [ 11,12], nanowires [ 13,14,15], and antidot \narrays [ 16] have been intensively investigated, mainly triggered by the discovery of \nspontaneous self -organization of magnetic FePt nanoparticles on a surface [ 11]. In all these \nfields, there is a rapid decrease of the relevant length scales of the magnetic structures down to the sub- 100 nm range. To illustrate this , a storage density of 1 Tbi ts/in\n2 would require a bit \nlength of around 25 nm. The underlying idea is to replace the comparatively large randomly -\noriented magnetic grains in conventional media with a nanoscopic region of a single magnetic \ndomain [17]. As the magnetic stability of individual particles scales with the material \nanisotropy constant and the particle volume , the use of FM FePt alloyed nanoparticles with \nvery high magnetic anisotropy (only in their chemically ordered face -centered tetragonal L1 0 \nphase) has been considered one of the promising routes towards future ultrahigh- density \nrecording -media applications [ 18,19]. In this case, the suppression of superparamagnetic \neffects at very small volumes might prove crucial to achieve stable and enhanced magnetic \nproperties at the smallest length scales. On the other hand, the use of ferrimagnetic \nnanoparticles with controlled coercivity near room temperature might be a more desirable \ncompromise to additionally achieve full control ove r the magnetic properties. 4 \n Recently, ferrimagnetic (FI) alloy systems with relatively low coercive fields , such as the \ncobalt (Co) - dysprosium (Dy) alloy DyCo 5, have been proposed as ideal candidates for \nmagnetic data storage and spintronic applications [20] owing to their flexibility in controlling \na tunable perpendicular exchange -bias effect at room temperature [ 21]. Here the Dy-Co alloy \nacts as a hard FI material with relatively low coercive field and well-defined perpendicular \nmagnetic anisotropy at room temperature . The system exhibits a spin -reorientation \ntemperature (TR) of ~350 K above which the uniaxial magnetic anisotropy rotates from out -\nof-plane to in -plane . Such a remarkable change of the anisotropy axis within a narrow \ntemperature window nea r room temperature renders this type of F I alloys one of the \ntechnologically relevant and promising materials for future spintronic applications. Therefore , \nparticularly taking into account that the magnetic properties of DyCo 5 nanostructures has \nremained unexplored, patterning of this type of FI alloys down to the nanoscale as well as \nunderstanding their magnetic properties at ultimate length scales are both of scientific and \ntechnological interest. \nAmong the different growth methods of magnetic nanostructure s, the anodization of \nalumina templates and the subsequent deposition of magnetic antidot array s have certain \nadvantages that overcome the unwanted effects of superparamagnetism. This technique \nemploys ultrahigh -density substrates th at are patterned down to the nanoscale as temp lates to \ngrow thin films on top [16,22]. In the case of nanoporous alumina membranes, nanoholes are \narranged in a fully -controllable hexagonal symmetry [ 23,24], and growing a thin film on top \nresults in an array of antidots [25]. The antidots act as well -ordered nonmagnetic inclusions \nwithin the magnetic thin film, promoting the creation of single nano- domain structures whose \nsizes can be manipulated depending on the anodization conditions. This is a much less \nexpensive technique compared to lithographic patterning methods. These nanostructures can \nbe promising candidates for a new generation of ultrahig h-density magnetic -storage media 5 \n mainly due to the absence of a superparamagnetic limit, as there are no isolated small \nmagnetic entities . This is due to the fact that the nanoholes introduce a locally distributed \nshape anisotropy [ 26], act themselves as pi nning centers for magnetic wall displacements, and \ntheir periodic distribution determines the whole magnetization process as well as the \nmagnetoresistance behavior [ 16]. Investigations of FI antidot arrays with tunable and \ncontrollable magnetic anisotropies might in fact provide superior functionalities that have not \nbeen addressed so far . \nIn this work, we present a nanoscale -patterned, hexagonally -ordered FI DyCo 5 antidot array \nwhere the local nanoscopic distribution of the magnetization at room temperature can be \ncontrolled via heati ng the sample slightly above its spin -reorientation temperature . We report \non controlling the magnetic anisotropy as a function of temperature and we compare the nano-\npatterned sample with its plain film counterpart. U sing x -ray magnetic circular dichroism \n(XMCD) in combination with photoelec tron emission microscopy (PEEM), we reveal that the \ntemperature- dependent changes of the magnetic anisotropy in the presence of antidots results \nin well- defined and stable magnetic nano -domain configurations modulating within ~45 nm at \nroom temperature . We discuss the underlying mechanisms giving rise to this behavior . Our \nresults pave the way for future applications of FI nanostructures in heat-assisted magnetic \nnano- devices . \nMETHODS \n We prepared antidot arrays of DyCo 5 by magnetron sputtering on top of hexagonally -\nordered alumina me mbranes , which were f abricated u sing a two -step anodization process [20, \n25]. For the present experiments, we produced membranes containing nanopores with a \ntypical diameter of ~68 nm and center -to-center inter-pore distances of ~105 nm [see Fig. \n1(a)] . A DyCo 5 extended film was grown on top of an un- patterned alumina substrate for 6 \n comparison purposes . The growth of the antidot arrays and the extended film was perform ed \nsimultaneously within the sputtering chamber, and their thickness was ~25 nm. The samples \nwere obtained by co-deposition from Co and Dy targets keeping the substrate temperature at \n~200°C. The co-deposition was done with a relative composition of 1:5 to ensure the spin-\nreorientation transition to be closest to room temperature. The samples were capped with a 2 \nnm thick Al layer to avoid oxidation upon transport in air. The magnetic properties of the \nextended films , their compensation and spin reorientati on temperatures, coercive fields, and \nmagnetic anisotropy axes were determined by element specific hysteresis loops at the Dy and \nCo absorption [ 21] edges u sing XMCD in transmission and scattering geometries [27]. \n High -resolution magnetic domain configurations around the nanoholes and across the spin-\nreorientation transition were obtained by XMCD -PEEM imaging [ 28]. Experiments were \nperformed at the UE49 -PGM a beamline of the synchrotron BESSY II . XMCD element -\nspecific images were obtained by tuning the synchrotron photon energy to the L 3 and M 5 \nresonances of Co (779.2 eV) and Dy (1293.9 eV), respectively . Each of the XMCD images \nwas calculated from a sequence of images taken with circular polarization (90% of circular \nphoton polarization) and alternating the light helicity. After normalization to a bright- field \nimage, the sequence was drift-corrected, and frames recorded at the same photon energy and \npolarization were averaged. T he difference of the two resulting images with opposite helicity, \ndivided by their sum , showed Co and Dy magnetic -domain contrasts, which represent the \nmagnetization vector pointing along the incidence direction of the x -ray be am. Due to the 16º \nshallow incidence angle of the x -rays on the sample, our XMCD- PEEM measurements are \nmainly sensitive to the in -plane magnetization. The samples were priorly magnetized by \napplying a magnetic field of ~0.5 T perpendicular to the surface, so that the XMCD images in the initial remanent magnetic state at room temperature exhibited nearly -zero magnetic \ncontrast. 7 \n RESULTS AND DISCUSSION \nWe investigated the formation of magn etic domains in FI DyCo 5 grown on extended \nsubstrates and on nanoporous membranes, controlling the sample tempe rature above and \nbelow the spin- reorientation transition temperature of ~ 350 K. Scanning electron microscopy \n(SEM) [Fig. 1(a)] and XPEEM images [ Fig. 1(b)] from the nanohole array show the \nhexagonal arrangement of antidots. Circles and lines on the SEM image mark several \npunctual defects and structural domains that occur during the self -assembling anodization \nprocess. The distance between the borde rs of two nanoholes, where the metallic alloy is \nsitting, is around 37 nm. The XPEEM image shows photoelectrons emitted from the material \ndeposited on the surroundings of the nanoholes as the bright intensity contrast, and dark \ncontrast corresponds to the nanohole locations. By varying the x -ray photon energy and \nsimultaneously rec ording XPEEM images, we obtain x -ray absorption spectra (XAS) for Co \n[Fig. 1(c)] and Dy [Fig. 1(d)], showing L 2,3 and M 4,5 absorption edges, respectively. \nAt a temperature of 385 K, just above T R = 350 K, the XMCD -PEEM magnetic contrast \nshown in Fig. 2 reveals clear in-plane ferrimagnetic ordering of Co and Dy elements, both \nbeing anti -ferromagnetically aligned with respect to each other in the imaged x -y surface \nplane. Very c learly, t he red (blue) regions of the Co XMCD images in Fig. 2 correspond to \nblue (red) regions of the Dy XMCD signal . Figure s 2(a -b) show the magnetic configuration \nfrom the antidot array sample and Fig s. 2(c -d) from its extended thin film counterpa rt. Re d \nand blue -color contrast correspond to magnetic domains pointing in the surface plane and \nprojected along the incident x -ray beam direction, with a maximum of ~10 % XMCD \nasymmetry. White (or faint) color contrast indicates eith er out -of-plane magnetization , \nmagnetization perpendicular to the incident beam direction or the zero net magnetization as it \nwould be expected from the nanohole positions. Comparing the XMCD -PEEM images of the \nantidot array to the ones of the extended film sample reveals clear differ ences in the size of 8 \n the magnetic domains; the antidot sample exhibits nanometer -sized domains separated by \nwhite color contrast, whereas the extended thin film sample exhibits domains in the order of \nseveral micrometers. This observation suggests that the antidots act as p inning centers for the \nmagnetization stabilizing magnetic nano -domains that are naturally separat ed from each other. \nBecau se the extended thin film lacks of such pinning centers, small nucleation domains which \ninitiate the magneti zation reversal have collapsed to form larger ones . \nIn the following, we present Co XMCD -PEEM images for three selected temperatures \nfrom a heating -and-cooling cycle; initial state at 300 K, high -temperature state at 470 K, and \nthe final state reached after cooling back to 300 K again, for both the extended thin film and \nantidot array samples. Figures 3(a) and 3(d) show the initial states for both samples, where nearly -zero local magnetization signal along the in- plane orientation is observed. The reason \nfor the vanishing magnetic contrast is the out -of-plane spin confi guration below T\nR. Upon \nheating to 470 K, which is well above the transition temperature, we clearly observe the \nappearance of in-plane oriented magnetic domains , as shown by the XMCD images in Fig. \n3(b) and 3(e). When we cool the samples back to room temperature and re -examine the \nresulting magnetic properties , the extended thin film sample re -establishes its original out-of-\nplane spin configuration, as expected [Fig. 3(c)] ; however, the antidot array exhibits a non-\nzero local magnetization as if the magnetic domains are mostly pinned to the in-plane spin \norientation even in the absence of any magneti c fields upon cooling [Fig. 3(f)]. To take a \ncloser look at the remanent local magnetization of the antidot array, we extract line profiles \nalong red and blue bits of magnetic domains as well as the real str ucture as seen by SEM and \nXPEEM. The results are summarized in Fig s. 3(g) and 3(h), where we show l ine profiles \nextracted along magnetic im ages for opposite magnetization directions (blue and red lines) . \nWe clearly observe magnetic information repeating every 45- 50 nm (between M ↑ and M=0 \nstates, and between M ↓ and M=0 states) , which correlates well with the difference between 9 \n the center -to-center inter -pore distance and the nanohole diameter . One bit of magnetic \ninformation being ~ 45 nm, one can estimate a data storage density of around 75 Gbits/inch2, \nwhich is in the same order of magnitude with the stora ge density of conventiona l hard -disk \ndrives. \nFigure 4 (a) summarizes key experimental results together with a schematic representation \nof the magnetic anisotropy Ku of the samples as a function of temperature. It shows various \nchannels of writing and storing data using a medium of DyCo 5 antidot array s. The easy axis \nof the magnetization rotates from out -of-plane to in- plane as the samples are heated above T R, \nand spins of the two sublattices of Co and Dy align in the sample plane. The change of the \neasy axis is outlined by the Co X MCD images from an antidot hexagonal unit cell and fr om \nthe extended thin film sample on the right -hand side of the schematic representation depicting \nthe in -plane anisotropy (K||) at high t emperatures. C ooling the extended thin film sample back \nto room t emperature leads to either one of the out -of-plane anisotropy configurations in the \nabsence of an external field , as seen in Fig. 3(c) . However, the antidot array sample has more \noptions upon cooling: it can decay into two in- plane and two out-of-plane ani sotropy \nconfigurations , as shown by the Co XMCD -PEEM images on the left -hand side of Fig. 4(a) . \nRegions within the antidot sample with out -of-plane magnetization show a faint color \ncontrast; therefore not much is seen in these images. Two regions of the sa mple with two \nhexagonal unit cells of antidots showing in- plane magnetization are also shown. We attribute \nthe occurrence of this magnetic multi- domain state to the presence of nanoholes , which act as \npinning centers for the magnetic anisotropy of the samp le. To illustrate this in more detail, in \nFigs. 4(b -d) we show the results of micromagnetic simulations within about one hexagonal \nunit cell obtained using the OOMMF package [ 29]. In Fig. 4(b) we show a SEM image from \nthe antidot array superimposed with blue circles around the nanoholes. The circles are used as \na mask to construct the input for the calculations. For simplicity, we simulate a Co layer with 10 \n the magnetic anisotropy oriented along the + y direction and an initial out- of-plane \nmagnetization. In Figs. 4(c) and 4(d) we show the calculated magnetization state for the \nantidot array and the extended film in the fully -relaxed state, respectively. Very clearly, the \nantidot ar ray in Fig. 4(c) exhibits a magnetic multi- domain configuration, while the extended \nfilm remains in a single -domain state in qualitative agreement with the experiments. The \ncalculations clearly reveal that the nanoholes pin the magnetization so that it follows circular \ncontours around them , leading to local changes in the magnetic anisotropy. These changes are \ncaused by the fact that every nanohole gives rise to six constrictions with its first neighbors, so that magnetic domains and domain walls are pinned or trapped by these constrictions. In \nthis way, every domain wall experiences a spatially -dependent dipolar interaction originating \nfrom a landscape of pinning potentials within the whole structure, resulting in local rotations of the magnetization and thus in magnetic multi- domain configurations . \nHaving observed these variety of magnetic domains in the experiment , one can consider \ncontrolling their occurrence by applying either in -plane or out -of-plane magnetic field pulses \n(H\n‖ or H┴) to selectively force the system to choose one of the four possibilities mentioned \nabove . The experimental results summarized in Fig. 4 (a) are therefore quite analogous to the \ncase where one applies a local magnetic field pulse during the cooling process to manipulate the mag netism of the nano- domains . It would be equally possible to employ this method for \nboth longitudinal and perpendicular recording as they follow the same basic principle. When \nthe energized write head passes by the medium, it leaves behind a magnetization p attern. For \nperpendicular recording, this magnetization pattern is ‘up’ and ‘down’ rather than ‘left’ and ‘right’ for longitudinal recording. \nContrary to the high anisotropy HAMR materials (among others: FePt, CoPt, SmCo\n5), FI \nantidots such as the ones st udied here would be a promising candidate that facilitates writin g \ndata by means of manipulating the magnetic easy and hard axes as a function of mild-11 \n temperature treatment. T he critical temperatu re of our magnetic system is not the Curie \ntemperature TC but the spin reorientation temperature TR, which is very close to room \ntemperature. Note that differently from this, high anisotropy HAMR materials need to be \nheated to temperatures close to their T C, which are around 750 -1000 K [ 30], to sufficiently \nreduce their high coercivity so that a moderate write field can induce a new magnetic state. In \nthis respect, there have been also attempts to reduce the high T C of the recording layers with \ndoping. In one example, Thiele et al. used Ni doping to reduce the T C of FePt; however, at the \nsame time the anisotropy decreases [ 31]. Assuming that in the near future plasmonics and \nnear-field optics will focus and transmit optical energy to spot sizes of around (25- 50 nm)2 \nand demagnetize the sample [ 19,30 ], lower phase -transition temperatures would still be more \nfeasible considering issues su ch as power consumption, excess -heat dissipation and risk of \ndamaging medium materials. \nSummarizing , we fabricated ferrimagnetic DyCo 5 antidot arrays for heat -assisted and bit -\npatterned magnetic recording. Our work de monstrates a novel functionality of ferrimagnetic \nbit-patterned materials that can be used to write and store magnetic data near room \ntemperature. The size of the nanoholes used in this work was around ~ 68 nm and their \nhexagonal lattice constant ~105 nm. W e have shown that the magnetization of the antidot \narray sample can be tuned from an out -of-plane to an in -plane spin configurat ion upon heating \nabove its spin- reorientation temperature of ~ 350 K. We have compared the magnetic \nproperties of the patterned s ample with the ones of its extended thin film counterpart and \nobserved remarkable differences in the size of the magnetic domains. Our results reveal the \nformation of small magnetic nano -domains in between antidots with well -defined and stable \nmagnetic dom ain configurations. Magnetic modulations of the nano- domains contrast are \nobserved to occur within ~45 nm for opposite magnetizations . We have attributed this \nbehavior to the change in the magnetic anisotropy caused by the nonmagnetic inclusions, 12 \n which at the same time act as pinning centers upon formation of multiple magnetic domains \nin agreement with our micromagnetic simulations . Finally, w e have discussed the \nimplications of our findings for heat -assisted magnetic recording using local magnetic probes. \nOur work demonstrates that ferrimagnetic antidot arrays exhibit tunable and controllable \nmagnetic traits with the potential to provide superior functionality for energy -assisted bit -\npatterned media near room temperature. \n \n \n 13 \n FIGURES \n \n \n \n \nFigure 1. (a) SEM and (b) XPEEM images of the hexagonal lattice antidot array of 68 nm pore size \nand 105 nm separation between pore centers . XAS spectra of (c) Co and (d) Dy elements at their L 2,3 \nand M 4,5 edges, as obtained from XPEEM energy scan s using linear horizont al polarized x -rays from \nthe field of view seen in (b). \n \n14 \n \n \n \n \nFigure 2. XMCD -PEEM images measured at 385 K at the Co L 3 and Dy M 5 edges from the antidot \narray (a- b) and the extended thin film sample (c -d) showing the ferrimagnetic ordering of the Co and \nDy magnetic moments via the corresponding blue -red XMCD contrast. Antidots stabilize nanometer -\nsized domains separate d from each other. \n \n \n15 \n \n \nFigure 3. XMCD -PEEM images measured at Co L 3 edge from the extended thin film sample (a -c) \nand from the antidot array sample (d -f) at the labeled temperatures. (g) Line profile along the structure \nof five neighbor ing antidots, XPEEM (black line) and SEM (dashed line). (h) Line profiles along \nXPEEM magnetic images for opposite magnetizations (blue and red lines) show magnetic information \nrepeating every 45 -50 nm (between M ↑ and M=0 states, and between M ↓ and M=0 states). \n16 \n \n \nFigure 4. (a) S chematic representation of the heat -assisted magnetic recording process and the spin \nreorientation transition as observed by XPEEM. A selection of images at the Co L3 edge summarizes \nthe experimental results. On the left, the room -temperature results correspond to different XMCD \nimages of the antidot array. (b-d) Results of micromagnetic simulations. In (b ), a SEM image from the \nantidot array superimposed with blue circles around the nanoholes is shown. The circles are used as a \nmask to construct the input for the calculations. The magnetic anisotropy is oriented along the + y \ndirection and the initial magnetization is out -of-plane. (c ) Calculate d multi -domain configuration of \nthe antidot array in the relaxed state. Small arrows depict the magnetization directions, blue and red \ncolors the projection of the magnetization along the y direction. (d) Same calculations as in (c ) for the \nextended film, revealing a single- domain configuration. \n17 \n AUTHOR INFORMATION \nCorresponding Author \n*E-mail: jaime.sanchez -barriga@helmholtz -berlin.de \nACKNOWLEDGMENT \nK. J. M and M. V gratefully acknowledge the support from MINECO under project \nMAT2013- 48054- C2-1-R. \n \nREFERENCES \n \n[1] S. S. P. Parkin , C. Kaiser, A. Panchula, P. M. Rice, B. Hughes, M. Samant, and S.- H. \nYang, Nature Mater. 3, 862 ( 2004). \n[2] M. N. Baibich, J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P. Eitenne, G. \nCreuzet, A. Friederich, and J. Chazelas, Phys. Rev. Lett. 61, 2472 (1988). \n[3] G. Binasch, P. Grünberg, F. Saurenbach, and W. Zinn, Phys. Rev. B 39, 4828 (1989). \n[4] S. S. P. Parkin, K. P. Roche, M. G. Samant, P. M. Rice, R. B. Beyers, R. E. \nScheuerlein, E. J. O’Sullivan, S. L. Brown, J. Bucchigano, D. W. Abraham, Y. Lu, M. Rooks, \nP. L. Trouilloud, R. A. Wanner, and W. J. Gallagher , J. Appl. Phys. 85, 5828 (1999). \n[5] R. L. Stamps, S. Breitkreutz, J. Åkerman, A . V. Chumak, Y . Otani, G . E. W. Bauer, \nJ.-U. Thiele, M. Bowen, S. A. Majetich, M. Kläui, I. L. Prejbeanu, B. Dieny, N. M. Dempsey, \nand B. Hillebrands, J. Phys. D: Appl. Phys. 47, 333001 ( 2014). \n[6] K. Bennemann, J. Phys .: Condens. Mat ter 22, 243201 (2010). \n[7] M. Molina -Ruiz, A. F. Lopeandi a, F. Pi, D. Givord, O. Bourgeois, and J. Rodr iguez -\nViejo , Phys. Rev. B 83, 140407(R) (2011). \n[8] D. Sander, J. Phys.: Condens. Matter 16, R603 ( 2004). \n[9] C. Desvaux, C. Amiens, P. Fejes, P. Renaud, M. Respaud, P. Lecante, E. Snoeck, and \nB. Chaudret, Nature Mater. 4, 750 ( 2005) . 18 \n \n[10] C. A. Ross, Annu. Rev. Mater. Res. 31, 203 ( 2001). \n[11] S. Sun, C. B. Murray, D. Weller, L. Folks, and A . Moser, Science 287, 1989 ( 2000) . \n[12] J. I. Martin, J. Nogues, K. Liu, J. L. Vicent, and I. K. Schuller , J. Magn. Magn. Mater. \n256, 449 ( 2003). \n[13] A. J. Yin, J. Li, W. Jian, A. J. Bennett, and J. M. Xu, Appl. Phys. Lett. 79, 1039 \n(2001). \n[14] J. Sánchez -Barriga, M. Lucas, F. Radu, E. Martin, M. Multigner, P. Marin, A. \nHernando, and G. Rivero , Phys. Rev. B 80, 184424 (2009). \n[15] J. Sánchez -Barriga, M. Lucas, G. Rivero, P. Marin, and A. Hernando, J. Magn. Magn. \nMater. 312, 99 ( 2007). \n[16] Z. L. Xiao, C. Y. Han, U. Welp, H. H. Wang, V. K. Vlasko- Vlasko, W. K. Kwok, D. \nJ. Millar, J. M. Hiller, R. E. Cook, G. A. Willing, and G. W. Crabtree, Appl. Phys. Lett. 81, \n2869 (2002). \n[17] T. R. Albrecht et al. , arXiv:1503.06664 (2015). \n[18] D. Weller , O. Mosendz , G. Parker , S. Pisana, and T. S. Santos , Phys. Status Solidi A \n210, 1245 (2013). \n[19] T. W. McDaniel, J. Appl. Phys. 112, 093920 (2012) . \n[20] F. Radu et al. , US Patent App. 14/383, 131 ( 2013) \n[21] F. Radu , R. Abrudan, I. Radu, D. Schmitz, and H. Zabel , Nat. Comm. 3, 715 ( 2012). \n[22] M. Vázquez , K. R. Pirota, D. Navas, A. Asenjo, M. Hernandez -Velez, P. Prieto, and J. \nM. Sanz, J. Magn. Magn. Mater. 320, 1978 ( 2008) . \n[23] H. Masuda and K. Fukuda, Science 268, 1466 (1995). 19 \n \n[24] A. Li, F. Müller, A. Birner, K. Nielsch, and U. Gösele, Adv. Mater. 11, 483 (1999). \n[25] K. Merazzo et al. , Phys. Rev. B 85, 184427 (2012); J . Appl. P hys. 109, 07B906 \n(2011) . \n[26] I. Ruiz -Feal, L. Lopez -Diaz, A. Hirohata, J. Rothman, C. M. Guertler, J. A. C. Bland, \nL. M. Garcia, J. M. Torres, J. Bartolome, F. Bartolome, M. Natali, D. Decanini, and Y. Chen, \nJ. Magn. Magn. Mater. 242- 245, 597 (2002) . \n[27] R. Abrudan, F. Brüssing, R Salikhov, J. Meermann , I. Radu , H. Ryll , F. Radu, and H. \nZabel , Rev. Sci. Instrum. 86, 63902 (2015) . \n[28] F. Kronast, J. Schlichting, F. Radu, S. K. Mishra , T. Noll, and H. A. Dürr , Surf. \nInterface Anal. 42, 1532 (2010) . \n[29] M. J. Donahue and D. G. Porter , OOMMF User’s Guide version 1.0 National Institute \nof Standards and Technology Inter Agency Report NISTIR 6376 ( 1999 ). \n \n[30] M. H. Kryder , E. C. Gage , T. W. McDaniel , W. A. Challener , R. E. Rottmayer , G. Ju, \nY.-T. Hsia , and M. F. Erden, Proc. IEEE 96, 1810 (2008) . \n[31] J.-U. Thiele, K. R. Coffey, M. F. Toney, J. A. Hedstrom, and A. J. Kellock, J. Appl. \nPhys. 91, 6595 (2002). " }, { "title": "1705.01335v1.Synthetic_ferrimagnet_spin_transfer_torque_oscillator__model_and_non_linear_properties.pdf", "content": "Synthetic ferrimagnet spin transfer torque oscillator: model and non-linear properties\nB. Lacoste,1M. Romera,2,∗U. Ebels,2and L. D. Buda-Prejbeanu2\n1International Iberian Nanotechnology Laboratory, Braga, Portugal\n2Univ. Grenoble Alpes, CEA, INAC-SPINTEC, CNRS, SPINTEC F-38000 Grenoble, France\n(Dated: Date of submission October 12, 2018)\nThe non-linear parameters of spin-torque oscillators based on a synthetic ferrimagnet free layer\n(two coupled layers) are computed. The analytical expressions are compared to macrospin simula-\ntions in the case of a synthetic ferrimagnet excited by a current spin-polarized by an external fixed\nlayer. It is shown that, of the two linear modes, acoustic and optical, only one is excited at a time,\nand therefore the self-sustained oscillations are similar to the dynamics of a single layer. However,\nthe non-linear parameters values can be controlled by the parameters of the synthetic ferrimagnet.\nWith a strong coupling between the two layers and asymmetric layers (different thicknesses), it is\ndemonstrated that the non-linear frequency shift can be reduced, which results in the reduction of\nthe linewidth of the power spectral density. For a particular applied field, the non-linear parameter\ncan even vanish; this corresponds to a transition between a red-shift and a blue-shift frequency\ndependence on the current and a linewidth reduction to the linearlinewidth value.\nKeywords: spin transfer torque, synthetic ferrimagnet, non-linear auto-oscillator\nSpin transfer torque oscillators (STOs) have promis-\ning applications as high frequency microwave generators.\nA typical STO nano-pillar is composed of two magnetic\nlayers separated by a metallic spacer or an isolating bar-\nrier. The magnetization of the first magnetic layer re-\nmains fixed in-plane or out-of-plane. It acts as a spin\npolarizer for the current flowing through the nano-pillar.\nThe magnetization of the second layer can be driven into\nself-sustained oscillations by an applied DC current due\nto spin transfer torque (STT)1–3. The oscillation of the\nfree layer magnetization gives rise to a variation of the\nresistance of the pillar, so that an alternative voltage ap-\npears at its boundaries. For a single domain free layer,\nthe generated microwave signal is typically in the GHz\nrange. However, the large linewidth, in the order of tens\nof MHz, is an obstacle for functional devices.\nIn order to improve the STO characteristics, an accu-\nrate and simple model describing the dynamics is funda-\nmental. For a single-layer (SL) STO, the general frame-\nwork of non-linear auto-oscillators (NLAO), proved to\nbe a particularly well adapted model4–7. Indeed, most\nof the features exhibited by experimental devices could\nbe explained within this framework, such as the field\nand current dependence of the frequency, the broadened\nlinewidth8, but also synchronization to an external signal\nor to other STOs9. More importantly, this model defines\na key parameter for understanding the STO behavior:\nthe non-linear amplitude-phase coupling parameter. By\nevaluating this non-linear parameter from the magnetic\nproperties of the layer, it was found that the linewidth of\nthe STO was reduced when applying a transverse field,\nfor instance5. However this model is confined to a SL free\nlayer and some recent works studied STO devices where\nthe free layer is composed of two coupled layers consti-\ntuting a synthetic ferromagnet (SyF)10,11. Allegedly, the\nadditional coupling energy would increase the magnetic\nstiffness and reduce the fluctuations. However, coupled\nsystems are also more complicated to understand and ageneral analytical model is necessary to explain and de-\nfinetheimportantparametersofitsdynamics. Typically,\nit would be useful to be able to calculate the non-linear\namplitude-phase coupling parameter of a SyF-STO.\nToanswerthisquestion, weproposetoextendthisframe-\nwork, the NLAO model, to describe the dynamics of two\ncoupled layers subjected to spin transfer torque. In order\nto treat the most general case, different coupling are in-\ncluded : the Ruderman-Kittel-Kasuya-Yosida (RKKY)\ninteraction, the dipolar coupling and the mutual STT.\nThe non-adiabatic STT (or field-like torque) is also in-\ncluded, although its effect was found to be negligible in\ntheparticularconfigurationsexaminedinthispaper. Itis\nfundamental in the dynamics of self-polarized STO12,13,\nthough.\nThe NLAO theory is based on a change of coordinates\nto complex variables to represent the magnetization dy-\nnamics of the layers. The phase and amplitude of the\ncomplexvariablesdescribethenon-lineardynamicsofthe\nauto-oscillator. The validity of this approach is limited\nto quasi-conservative trajectories, for which the energy is\nalmost constant, and to small oscillation amplitudes. Us-\ning common diagonalization techniques, the conservative\npart of the magnetization equation of motion is simpli-\nfied to two terms : a linear and a non-linear contribution.\nThe dissipative part, which is supposed to be small com-\npared to the conservative part, defines the equilibrium\nenergy of the auto-oscillator by balancing the negative\nGilbert damping and the positive STT.\nThe first and second parts of this work describe the steps\nto extract the auto-oscillator equation for two coupled\nSyF layers in the macrospin approximation. In the third\nand fourth part, we describe the dynamics of the SyF-\nSTO defined by two coupled equations, so the STO can\nbe described by two modes. However, only one of them\nis usually excited into steady-state at a time, so the SyF-\nSTO is equivalent to a single-layer (SL) STO. This is an\nimportant result of this paper. The parameters of thearXiv:1705.01335v1 [cond-mat.other] 3 May 20172\nsingle-mode SyF-STO are computed, especially the non-\nlinear parameter, which is responsible for the frequency\ntunability, the large linewidth and the synchronization\nbandwidth. Another important result of this paper is\nthe link between the vanishing of the non-linear parame-\nter and the transition between a redshift and a blueshift\nregime. Finally in the fifth part, we study how to de-\ncrease the linewidth of a SyF-STO by changing the cou-\npling strength and the thickness of the layers.\nI. DESCRIPTION OF THE SYSTEM\nA. Landau-Lifshitz-Gilbert-Slonczewski equation\nWe consider the system in Figure 1 of two magnetic\nlayers, labeled 1 and 2 constituting a synthetic ferromag-\nnet (SyF). The total free energy Eholds the demagnetiz-\ning energy, the uniaxial anisotropy energy and the Zee-\nman energy (including an exchange energy) of both lay-\ners, plus a conservative coupling term between the two\nlayers, consisting of an RKKY interaction coupling and\nthe dipolar coupling :\nE=µ0\n2V1M1Hd1(m1·uz)2−µ0\n2V1M1Hk1(m1·ux)2\n+µ0\n2V2M2Hd2(m2·uz)2−µ0\n2V2M2Hk2(m2·ux)2\n−˜Dxm1xm2x−˜Dym1ym2y−˜Dzm1zm2z\n−µ0V1M1(Hx+Hex 1)(m1·ux)\n−µ0V2M2(Hx+Hex 2)(m2·ux) (1)\nHereµ0is the permeability of free space. V1=t1S\nandV2=t2Sare the volumes of the layers, with\nthicknesses t1andt2and surface S.M1andM2are\nthe saturation magnetizations of the layers, Hd1,Hd2\ntheir demagnetizing fields (supposed positive), Hk1,Hk2\nthe uniaxial anisotropy fields. For each layer labeled\nbyi= (1,2), we define the demagnetizing coefficients\n(Ni\nxx,Ni\nyy,Ni\nzz), and the interface anisotropy constant\nKSi, so thatHdi= (Ni\nzz−Ni\nyy)Mi−2KSi/(µ0Miti)\nandHki= (Ni\nyy−Ni\nxx)Mi.Hex1,Hex2are the exchange\nfields acting on each layer (for instance from a coupling\nwith a fixed anti-ferromagnet), and Hxthe applied field\nalong the easy axis.\nThe coefficients ˜Dx,˜Dyand ˜Dzaccount for the (conser-\nvative) coupling between the two layers. They include\nRKKY interaction term and the dipolar coupling, such\nthat, fori= (x,y,z ),˜Di=SJRKKY +Di, where\ntheDiare the dipolar coupling energy coefficients in\nmacrospin. Note that a negative JRKKYcorresponds to\nan anti-ferromagnetic coupling between the layers. Two\nadjacent layers give rise to negative DxandDy, and\npositiveDz.\nThe layers are also subject to a spin transfer torque\n(STT) due to a current flowing perpendicular to the lay-\ners. A positive current corresponds to electrons flowing\nFigure 1. Schematics of the synthetic ferrimagnet (SyF) as a\nfree layer with a fixed in-plane magnetized reference layer.\nfrom layer 2 towards layer 1, and then to the reference\nlayer. Thus, layer 1 is subjected to the STT from the\nreference layer and to the STT from layer 2 (with a neg-\native factor because layer 1 receives reflected electrons\nfrom layer 2). Layer 2 is subjected to the STT from the\nreferencelayerandfromlayer1(becauseofelectronsthat\nwere spin polarized after passing through layer 1). These\nspin torques acting on the two layers are modeled by two\nspin torque potentials, for layer 1 and 2 respectively, P1\nandP214:\nP1=−~\n2|e|Iη1m1·ux+~\n2|e|Iη21m1·m2\nP2=−~\n2|e|Iη2m2·ux−~\n2|e|Iη12m1·m2\nIis the current flowing through the layers. η1(resp.η2)\nis the effective spin-polarization of the current in layer\n1 (2) due to the fixed in-plane polarizer positioned be-\nfore layer 1 according to the direction of the current. η12\n(resp.η21) is the effective spin-polarization of the current\nin layer 2 (1) due to layer 1 (2).\nMoreover, the two layers are subjected to perpendicu-\nlar (or field-like) spin transfer torque (pSTT), from the\nreference layer and from the other layer. The pSTT is\nmodeled by two potentials, similar to the spin torque po-\ntentials defined above :\n˜P1=−~\n2|e|Iβ1m1·ux+~\n2|e|Iβ21m1·m2\n˜P2=−~\n2|e|Iβ2m2·ux−~\n2|e|Iβ12m1·m2\nThe equation of motion is given by the Landau-\nLifshitz-Gilbert-Slonczewski (LLGS) equation. In this\nform, the damping is defined with respect to the time-\nderivative of the magnetization vector; after moving all\nthe time-derivatives on the left-hand-side, the LLGS3\nwrites :\nµ0V1M1dm1\ndt=γ0m1×∂E\n∂m1+γ0m1×∂˜P1\n∂m1\n+γ0m1×/parenleftbigg\nm1×∂\n∂m1(P1−α1E)/parenrightbigg\nµ0V2M2dm2\ndt=γ0m2×∂E\n∂m2+γ0m2×∂˜P2\n∂m2\n+γ0m2×/parenleftbigg\nm2×∂\n∂m2(P2−α2E)/parenrightbigg\nThe Gilbert damping coefficients of the two layers are\ngiven byα1andα2. The correction to the gyromagnetic\nratio due to the damping coefficient has been neglected,\nsoγ0=µ0γwhereγis the gyro-magnetic ratio.\nAccording to the form of the LLGS equation used in\nthis paper, the coefficients βj(j= (1,2,12,21)) of the\nfield-like torques can contain a term proportional to the\ncoefficients ηjfrom the damping-like STT and to the\nGilbert damping constants of the two layers, that we call\npseudo-field-like torque. Namely β1=α1η1,β2=α2η2,\nβ12=α2η12andβ21=α1η21. Such additional terms\nwould be coming from the transformation of the STT\nfrom the Gilbert-form of the LLGS equation to the\nLandau-form.\nBy writing the LLGS equation in this form, the\nfree energy part, which is common to both layers, is\nseparated from the rest. This will allow to use a similar\nformalism as for the description of a single layer in\nprevious publications4,5.\nIn order to simplify the notations, we introduce the\nlayer asymmetry β, the geometrical mean magnetic vol-\numeMand the following normalized hamiltonian and\npotentials :\nβ=/radicalbigg\nM2t2\nM1t1M=µ0S/radicalbig\nM1t1M2t2\nH=γ0E\n2M∆1=γ0˜P1\n2M∆2=γ0˜P2\n2M\nΓ1=α1H−γ0P1\n2MΓ2=α2H−γ0P2\n2M\nIn the following, dotted variables represent their time\nderivative. Therefore the LLGS equation rewrites :\n1\n2β˙m1=m1×∂H\n∂m1+m1×∂∆1\n∂m1\n+m1×/parenleftbigg\nm1×∂Γ1\n∂m1/parenrightbigg\n(2)\nβ\n2˙m2=m2×∂H\n∂m2+m2×∂∆2\n∂m2\n+m2×/parenleftbigg\nm2×∂Γ2\n∂m2/parenrightbiggWe decompose the right-hand-side of the LLGS equation\nin two parts that will be treated separately: (i) the con-\nservative hamiltonian terms (simply called conservative\nin the following) that are composed of the first terms\non the right-hand-side and depend only on H. (ii) the\nconservative non-hamiltonian terms and the dissipative\nterms (simply called dissipative in the following because\nthe dissipative terms play a more important role) that\nare composed of the other two terms (respectively) on\nthe right-hand-side.\nIn general, the damping constants α1,α2are considered\nto be small ( <0.1) and the applied current is reasonably\nsmall, so the conservative part is larger than the dissipa-\ntive part. The two different orders of magnitude further\nsupport the distinction made between the two parts.\nB. Numerical parameters\nThe results from the extended NLAO model will be\ncompared to macrospin LLGS simulations. The case of\nan asymmetric SyF shows interesting properties, espe-\ncially in terms of linewidth reduction. As all cases can-\nnot be reproduced here, we focus on a SyF with thickness\nasymmetry between the two layers. However, asymme-\ntry can also be introduced by submitting one layer to an\nexchange field or by reducing the effective demagnetiz-\ning field of one of the layers with perpendicular interface\nanisotropy.\nFor the inter-layer coupling, two regimes are considered,\nsmall coupling JRKKY =−2×10−4J/m2and large cou-\nplingJRKKY =−5×10−4J/m2. The dipolar coupling\nis neglected in the macrospin simulations. This is sup-\nported by the fact that, in the macrospin approximation\nandinnano-pillarswithcircularcross-section,thedipolar\ncoupling is an antiferromagnetic coupling in the in-plane\ndirections ( xandydirections in our convention) and a\nferromagnetic coupling in the normal direction ( zdirec-\ntion). Because of the high demagnetizing field in thin\nlayers, the trajectories have a small out-of-plane compo-\nnent, so contribution from the dipolar coupling is com-\nparable to a low RKKY antiferromagnetic coupling. For\nthe layer thicknesses considered, the dipolar field is lower\nthan the RKKY coupling field, so the dipolar coupling is\nsimply neglected.\nThe rest of the parameters are defined in Table I.\nAccording to the value of the area Sof the pillars, cur-\nrents expressed in mA correspond to current densities of\n1011A/m2.\nThe current is considered to be unpolarized after going\nthrough the first layer, so η2= 0. However the same\nqualitative results were obtained15if we suppose that\nη2=±η1.\nThe conservative part is the most important to de-\nscribe the self-sustained oscillations because the trajec-\ntories of the self-sustained oscillations are close to the\nconstant energy trajectories. For this reason, a change of\nvariables that describes accurately the conservative part4\nIdentical properties Value\nMs1,Ms2 1×106A/m\nHd1,Hd2 0.9×106A/m\nHk1,Hk2 10×103A/m\nHex1,Hex2 0\nα1,α2 0.02\nS 10−14m2\nη21,η12 0\nβ1,β2,β12,β21 0\nDifferent properties Values\nt1andt21.8 and 2.2 nm\nη1andη2 0.5 and 0\nTable I. Properties of the magnetic layers.\nand treats the dissipative part as a small perturbation is\nadapted to describe the dynamics of the STO. This will\nbe developed in the next part.\nII. TRANSFORMATION TO COMPLEX\nVARIABLES\nA. Complex variables: conservative part\nWe intend to rewrite the LLGS equation in complex\nform representing the evolution of two modes a1anda2.\nLeta= (a1,a2)be a 2-dimensional complex vector. The\ngoal is to write the conservative part of the LLGS equa-\ntion in the form :\n˙a=−i∂H\n∂a†(3)\nThe elements of the basis, a1anda2, represent uni-\nform modes around the equilibrium position, with com-\nplex conjugates a†= (a†\n1,a†\n2). In the following, we focus\nonly on the modes around the parallel equilibrium state\n(or antiparallel depending on the sign of the RKKY cou-\npling constant and the dipolar coupling), i.e. the syn-\nthetic ferrimagnet (SyF) is in the plateau region. The\nequilibrium position is represented by :\nmeq\n1x=mux meq\n2x=mnux\nHeren= sign( ˜Dx)reflects the ferromagnetic or anti-\nferromagnetic type of coupling between the two lay-\ners:n= +1ferromagnetic coupling, n=−1anti-ferromagnetic coupling. The direction of layer 1 rela-\ntively to the fixed reference layer is given by m:m= +1\nfor a parallel (P) orientation, m=−1for an antiparallel\n(AP) orientation. The initial state is then defined by a P\nor AP configuration (with respect to the reference layer)\nand a ferromagnetic or anti-ferromagnetic coupling be-\ntween layer 1 and layer 2.\nWe proceed to a change of coordinate system so that\nthe equilibrium magnetizations have the same definitions\nfor all the layers. They are defined by meq\ni=ui\nζfor\ni= (1,2):\nu1\nζ=mux u2\nζ=mnux\nu1\nξ=muy u2\nξ=mnuy\nu1\nη=uz u2\nη=uz\nThe expressions of a1anda2with respect to the local\nmagnetization coordinates have to be chosen adequately\nso that the conservative part of LLGS in this new system\nof coordinates take the hamiltonian form of Eq. (3). For\nthat we set :\na1=1√βm1ξ−im1η/radicalbig\n2(1 +m1ζ)(4)\na2=/radicalbig\nβm2ξ−im2η/radicalbig\n2(1 +m2ζ)(5)\nNotice that there are other choices of (H,a1,a2)that al-\nlows to rewrite the LLGS equation in the hamiltonian\nform of Eq. (3), notably by multiplying a1anda2by the\nsame constant term CandHbyC2. The quadratic part\n(as it will be defined later) of the Hamiltonian would re-\nmain unchanged by changing this factor, but the quartic\n(and the other orders) part would be affected. Hence, it\nis not possible to compare coefficients of quartic or higher\norder for different geometries, as their definition depends\non the arbitrary choice of the constant C. Instead, nor-\nmalized coefficients should be compared.\nThe expression of Hwith respect to the new variables\n(a1,a2)and their complex conjugates (a†\n1,a†\n2)can be di-\nvided asH=H2+H4by dropping the constant term\nand neglecting higher order hamiltonian terms. In terms\nof the complex variables, H2is the quadratic part and\nH4is the quartic part.\nH2=A1a1a†\n1+A2a2a†\n2+1\n2/parenleftbig\nB1a2\n1+B2a2\n2+c.c./parenrightbig\n+/parenleftbig\nC12a1a2+D12a1a†\n2+c.c./parenrightbig\nH4=U1a2\n1a†2\n1+U2a2\n2a†2\n2+W12a1a2a†\n1a†\n2\n+/parenleftbig\nV1a3\n1a†\n1+V2a3\n2a†\n2+c.c./parenrightbig\n+/parenleftbig\nY12a2\n1a†\n1a2+Y21a1a2\n2a†\n2+c.c./parenrightbig\n+/parenleftbig\nZ12a1a†2\n1a2+Z21a†\n1a2\n2a†\n2+c.c./parenrightbig\nWe introduce new parameters that correspond to the5\ncharacteristic frequencies :\nω1\nk=γ0Hk1,ω2\nk=γ0Hk2,ω1\nd=γ0Hd1,ω2\nd=γ0Hd2,\nω1\na=γ0m(Hx+Hex1),ω2\na=γ0nm(Hx+Hex2),\nω0\nc=γ0\nMn˜Dx,ω−\nc=γ0\nMn˜Dy−˜Dz\n2,ω+\nc=γ0\nMn˜Dy+˜Dz\n2.\nUsing these notations, the coefficients of the hamiltonian\nare given by :\nA1=ω1\nk+ω1\nd\n2+ω1\na+βω0\nc,A2=ω2\nk+ω2\nd\n2+ω2\na+ω0\nc\nβ,\nB1=−ω1\nd\n2,B2=−ω2\nd\n2,C12=−ω−\nc,D12=−ω+\nc,\nU1=−βω1\nk−β\n2ω1\nd,U2=−ω2\nk\nβ−ω2\nd\n2β,\nW12=−2ω0\nc,V1=β\n4ω1\nd,V2=ω2\nd\n4β,\nY12=β\n2ω−\nc,Y21=ω−\nc\n2β,Z12=β\n2ω+\nc,Z21=ω+\nc\n2β.\nIn matrix form,H2rewrites :\nH2=1\n2/parenleftbig\na†\n1a†\n2a1a2/parenrightbig\nA1D12B1C12\nD12A2C12B2\nB1C12A1D12\nC12B2D12A2\n\na1\na2\na†\n1\na†\n2\n\nThe notation xis used for the complex conjugate of\nx, to distinguish scalar coefficients from the magnetiza-\ntion complex variables (a1,a2)with complex conjugates\n(a†\n1,a†\n2). As for a SL oscillator, it is possible to diago-\nnalize the quadratic part H2of the hamiltonian16. In\nfact it is possible to do so for any number of layers, al-\nthough it becomes difficult to find analytical expressions\nformorethantwolayers. Thenewcomplexbasisiscalled\n(bop,bac), with :\n\na1\na2\na†\n1\na†\n2\n=Tab\nbop\nbac\nb†\nop\nb†\nac\n(6)\nTab=\nuop\n1uac\n1vop\n1vac\n1\nuop\n2uac\n2vop\n2vac\n2\nvop\n1vac\n1uop\n1uac\n1\nvop\n2vac\n2uop\n2uac\n2\n\nWe note/hatwideIthe 4x4 block matrix /hatwideI=/parenleftbigg\nI20\n0−I2/parenrightbigg\nwithI2\nthe2×2unity matrix. T†\nabis the transpose conjugate of\nTab. It verifies :\nT−1\nab=/hatwideIT†\nab/hatwideIIn the new basis, H2takes the simple form :\nH2=ωopbopb†\nop+ωacbacb†\nac\nThe complex variables (bop,bac)are eigenvectors of the\nlinear hamiltonian. They correspond to the two linear\nmodes of the SyF-STO: optical and acoustic. This base\nof eigenvectors is then used to express the non-linear part\nof the hamiltonian.\nThe expressions of ωop,ωacand of the coefficients of\nthe matrix Tabcome from diagonalizing the matrix ˜H2:\n˜H2=\nA1D12B1C12\nD12A2C12B2\n−B1−C12−A1−D12\n−C12−B2−D12−A2\n\nThe expression of ˜H2is for the general case, for any di-\nrection of the equilibrium magnetizations. In the con-\nfiguration studied here, with equilibrium configurations\nand applied fields along the easy axis, all the coefficients\nare real. We take this assumption in the following.\nFrom computing the eigenvalues of ˜H2, the following val-\nues are obtained for ωop/ac:\nω2\nav= (A2\n1+A2\n2)/2.−(B2\n1+B2\n2)/2.+D2\n12−C2\n12\n∆ =/parenleftBig\nA2\n1−A2\n2−B2\n1+B2\n2/parenrightBig2\n+ 4/parenleftBig\nC12(B1+B2)−D 12(A1+A2)/parenrightBig2\n−4/parenleftBig\nC12(A1−A 2)−D 12(B1−B2)/parenrightBig2\nω2\nop/ac =ω2\nav±√\n∆\n2(7)\nThefrequencies ωopandωaccorrespondtothetwomodes\noptical and acoustic that are observed in ferromagnetic\nresonance (FMR) experiments with SyFs. By definition,\nthe optical mode corresponds to the mode with the high-\nest frequency. The expressions of the two mode frequen-\ncies are in agreement with the expressions found in the\nliterature14,17.\nThe eigenvectors of ˜H2, which correspond to the\ncolumns of the matrix Tab, have complicated expressions.\nHowever, due to normalization conditions, they can be\nexpressed by 6 angles. For the two labels j= (op,ac),\nthe elements of the matrix Tabare given by :\nuj\n1= coshθjcosφj\nuj\n2=−coshθjsinφj\nvj\n1=−sinhθjcosψj\nvj\n2= sinhθjsinψj\nThe details about the coefficients are given in Ap-\npendix A.\nThe angles φjandψjare related to the coupling6\nbetween the two layers. In fact, if the coupling vanishes\n(C12=D12= 0), these angles vanish for one mode,\nsay the acoustic mode, φac=ψac= 0, whereas for the\nother mode, φop=ψop=π/2. So the optical mode bop\ndepends only on the layer 2 complex variable a2and the\nacoustic mode bacon the layer 1 and a1.\nThe angles θjcorrespond to the mixing between the\ndiagonal termsA1,A2and the off-diagonal terms B1,\nB2, by analogy to the transformation coefficients for a\nsingle layer.\nHowever, it is not possible to obtain an exact diago-\nnalization of the quartic part H4of the hamiltonian but\nnon-canonical transformations provide good approxima-\ntions. We distinguish the resonant terms, for which the\noverall phase vanishes, like bopb†\nop, from the non-resonant\n(or off-diagonal) terms, for which the overall phase varies\nwith time, like bopb†\nac.\nBecause in this configuration, all along the easy axis,\nthere is no cubic term in the Hamiltonian, the (non-\ncanonical) transformation to remove the conservative\nnon-resonant terms18does not affect the value of the di-\nagonal quartic terms. Equivalently, we then assume that\nthe off-diagonal terms of the quartic term are negligi-\nble. However this assumption is valid only if the mode\nfrequencies ωopandωacare large compared to the off-\ndiagonal terms. Concretely, when applying an external\nfield that is comparable to the spin-flop field, the acous-\ntic mode frequency almost vanishes and the previous as-\nsumption is no longer valid. In this case the dynamics is\nmore complicated because the conservative non-resonant\nterms become important. We therefore limit the major\ndiscussion to the field range below the spin-flop field.\nNeglecting the non-resonant terms, the quartic part has\nthe simple expression :\nH4=Nop\n2b2\nopb†\nop2+Nac\n2b2\nacb†\nac2+Tbopb†\nopbacb†\nac\nWhereNac(resp.Nop) is the acoustic (optical) non-\nlinearfrequencyshiftcoefficientand Tisthemixed-mode\nnon-linear frequency shift coefficient. They are all real.\nAll these coefficients come from the conservative part of\nthe LLGS equation, they depend on the demagnetizing\nfields of the layers, applied field and coupling energy.\nHowever, they are independent of the damping coeffi-\ncients of the layers and of the applied current.\nB. Complex variables: dissipative part\nWenowfocusonthedissipativepartoftheLLGSequa-\ntion. After the transformation to the complex variables\na1,a2, the LLGS equation writes :\n˙a=−i∂H\n∂a†−Fa (8)\nWhere Fa= (Fa1,Fa2)is a vector with two complex\ncomponents. The two dissipative complex componentsFa1,Fa2are truncated to contain only linear and cubic\nterms ina1,a2,a†\n1anda†\n2. The polynomial coefficients\nare noted with 4 indices (k,l,m,n ), so that :\nFai=/summationdisplay\nk,l,m,nfk,l,m,n\naia1ka2la†\n1ma†\n2nfori= 1,2\nTheexpressionsofthecoefficientsofthedissipativeterms\nare given in the Appendix B. Using the linear transform\nwiththematrix Tab,similarcoefficientsforthe b-variables\nare obtained, with b= (bop,bac):\n˙b=−i∂H\n∂b†−Fb\nFb= (Fbop,Fbac)and fori=op, ac :\nFbi=/summationdisplay\nk,l,m,nfk,l,m,n\nbibopkbaclbop†mbac†n\nSo that :\n/parenleftbiggFb\nF†\nb/parenrightbigg\n=T−1\nab·/parenleftbiggFa\nF†\na/parenrightbigg\nWhere FaandF†\naare expressed in terms of b-variables\nusing the transform of equation (6).\nAll these fk,l,m,n\nbicoefficients in the b-coordinates\nare complex in general. However, if the coefficients\nof the conservative terms are real ( B1,B2,C12,D12,\nV1,V2, etc...), only the field-like torque contributes to\nthe imaginary part. For the magnetic configuration\nstudied in this paper, the applied field is aligned with\nthe magnetization, so the conservative coefficients are\nreal. The field-like torque is also set to zero, so the\ndissipative coefficients are real. In any case, the real part\nis the really important part, as it defines the power as it\nwill be shown in the next section. The imaginary part\nonly gives a contribution to the phase equation and it is\nnegligible compared to the contribution from the con-\nservative part in the configuration studied in this paper,\nwith an external polarizer. Without external polarizer,\nbut taking into account the mutual spin-torque in a\nself-polarizer structure, the contribution of the field-like\ntorque is non-negligible as shown in reference 19.\nOf all the dissipative terms, the most important are\nthe resonant terms, i.e. the terms that are similar to the\nresonant terms from the conservative part. Taking into\naccount only these resonant terms, the dissipative part\nreduces to :\nFbop=bop/parenleftBig\nγop+Qopbopb†\nop+Ropbacb†\nac/parenrightBig\nFbac=bac/parenleftBig\nγac+Qacbacb†\nac+Racbopb†\nop/parenrightBig\n(9)\nγac(resp.γop) is the acoustic (optical) linear relaxation\nrate.Qac(Qop) is the acoustic (optical) non-linear\nrelaxation rate coefficient. Rac(Rop) is the coefficient of\nthe acoustic (optical) non-linear mode mixing relaxation7\nrate.\nBecause of the linear dependence of the STT amplitude\nwith respect to the applied current hypothesized in this\npaper, these coefficients depend linearly on the applied\ncurrent. The linear coefficients, γopandγac, are positive\nfor zero current, in agreement with the fact that the\nGilbert damping is a relaxation to the minimum energy\nconfiguration. They decrease with the current if the\ncurrent is applied in the direction that destabilizes the\nmagnetization. The dissipative coefficients also depend\non the demagnetizing fields and coupling energy, like the\nconservative coefficients.\nThe analytical expressions of the coefficients are very\nlengthy and therefore they are not presented here in de-\ntail. Instead, for each value of field and current, the co-\nefficients are calculated numerically through the various\ntransformations, using the materials parameters given in\nsection IB. The variation of the different coefficients with\nfield and current are given in section III, where the cou-\npled complex equations are solved.\nIII. DYNAMICS WITH RESONANT TERMS\nONLY\nIn order to illustrate some of the basic features of the\ncoupled system, in a first approximation only the reso-\nnant terms, i.e.H2,H4and Eq. 9 are considered for the\ntime evolution of bopandbac:\n˙bop=−ibop(ωop+Noppop+Tpac)\n−bop(γop+Qoppop+Roppac)\n˙bac=−ibac(ωac+Nacpac+Tpop)\n−bac(γac+Qacpac+Racpop)(10)\nWherepop=bopb†\nopandpac=bacb†\nacare the powers of\nthe two modes. All the coefficients are supposed to be\nreal.\nWe notice that the dynamics of the coupled system does\nnot reduce to two independent oscillator equations. Even\nif the two modes are decoupled in the linear regime\n(pop,pac/lessmuch1), the acoustic and optical modes are cou-\npled through the non-linear coefficients.\nIntroducing the phases φop,φacof the two modes, let’s\ndefine :\nbop=√pope−iφop\nbac=√pace−iφac\nUsing the definitions of bopandbac, one can derive\nseparate equations for the power and the phase. These\nwill be discussed in the next sections. It is reminded\nthat for a single layer the equivalent analytical equations\nyield as a stationary solution a constant oscillation\npower (cancellation of the dissipative part). In the next\nsection it is shown that the coupled Eq. 10 can reduce\nto a single mode equation under specific conditions.\nFigure 2. Linear and non-linear dissipative coefficients versus\napplied current IforHx=−40kA/m and JRKKY =−5×\n10−4J/m2. (a) Optical coefficients, (b) acoustic coefficients.\nAll values are divided by 2πto be in units of Hz and not in\nrad/s.\nFor this we start discussing the solutions to the power\nequations.\nA. Power equations\nTheequationsoftimeevolutionofthepowerandphase\nare derived from the complex equations 10. The equa-\ntions of evolution of the powers of both modes are given\nby the generalized Lotka-Volterra (LV) equations20:\n˙pop=−2pop(γop+Qoppop+Roppac)\n˙pac=−2pac(γac+Qacpac+Racpop)(11)\nLotka-Volterra systems are well known for modeling the\nevolution of predator-prey populations. We define the\nsingle-mode equilibrium powers ¯popand¯pacas :\n¯pop=−γop\nQop¯pac=−γac\nQac(12)\nTheeffective linear coefficients are defined by :\ndop=γop+ ¯pacRopdac=γac+ ¯popRac\nAnd the inter-mode mixing coefficient ∆is defined by :\n∆ = 1−RopRac\nQopQac\nThe convergence to equilibrium for the LV system is\ndescribed in reference 21 and references 22, 23 provide a\nclassification with state diagrams. The two-modes sys-\ntem has four equilibriums, their conditions for existence\nand stability are defined by :8\n•P0= (0,0)ifγop>0andγac>0:\nNo mode is excited, this is the subcritical regime\nwith only damped modes.\n•Pop= (¯pop,0)ifγop<0,Qop>0anddac>0:\nOnly the optical mode is excited and the acoustic\nmode vanishes.\n•Pac= (0,¯pac)ifγac<0,Qac>0anddop>0:\nOnly the acoustic mode is excited and the optical\nmode vanishes.\n•P∗= (p∗\nop,p∗\nac)ifdopQac<0,dacQop<0,\nQopQac∆>0and(dop+dac)/∆<0:\nThe system converges to a mixed-mode equilibrium\nwhere both modes have a finite power given by:\np∗\nop=−dop\nQop∆p∗\nac=−dac\nQac∆(13)\nNotice thatP0andP∗are compatible, they can be\nstable local equilibriums at the same time, but they are\nincompatible with PopandPac. And reciprocally, Pop\nandPaccan be stable at the same time, but not at the\nsame time asP0andP∗.\nGiven the specific conditions are fulfilled, each equilib-\nrium is defined and locally stable. However, the global\nconvergence to this equilibrium depends on the initial\nconditions, if they are in the basin of convergence of this\nequilibrium. For instance, PopandPaccan be stable\nat the same time, it depends on the initial conditions if\nthe system converges to one or the other equilibrium,\nor even if it diverges (which corresponds to a switching\nof one or both layers). See Fig. 4 in reference 20 for a\nphase portrait of pacversuspop— notedn1andn2.\nThe coefficients of Eq. (11) are plotted in figure 2\nversus applied current Ifor the macrospin parame-\nters defined previously and for Hx=−40kA/m and\nJRKKY =−5×10−4J/m2. Two threshold currents\nfor the modes excitations, Iac\ncandIop\nc, are defined by\nthe vanishing of the linear coefficients γopandγac,\nrespectively. For this particular set of parameters, the\nacoustic threshold current Iac\ncis lower than the optical\nthreshold current Iop\nc. Therefore the critical current Ic\ncorresponds to the acoustic threshold current, which is\nIc= 3.4mA in this particular case. Above the critical\ncurrentIc, the acoustic mode is excited, and because Qac\nis positive (not shown in Figure 2 above 2mAQacin-\ncreases linearly), the power converges to the equilibrium\nacoustic power; the optical mode remains zero. Above\nthe optical threshold current, the equilibrium acoustic\npower still exists and it is stable, because dop>0(not\nshown on the figures). However, Qopis negative, so\nno equilibrium optical power is defined and the optical\nmode may diverge. Therefore, the final state depends\non the initial conditions : if the acoustic power is close\nto the equilibrium ¯pacand the optical power is close to\n0, the system converges to the powers {0; ¯pac}; if theoptical mode diverges faster than the acoustic mode\nconverges to its equilibrium value, the whole system\nwill diverge, which corresponds to a reversal of the layers.\nHaving defined the equilibrium powers, the oscillation\nfrequency is given by the phase equations that will be\nanalyzed in the next section.\nB. Phase equations\nThe corresponding phase equations of Eq. 10 including\nonly resonant terms are :\n˙φop=ωop+Noppop+Tpac\n˙φac=ωac+Nacpac+Tpop (14)\nWe notice that the phase velocities ˙φopand ˙φacof the\ntwo modes are constant if the powers are at equilibrium\n(˙pop= ˙pac= 0). Moreover, the phase of each mode\ndepends not only on its own power, but also on the power\nof the other mode through the non-linear phase mixing\nT. But both phases are independent of each other : each\nmode oscillates at its own constant frequency. Note that\nthis is true only if the non-resonant terms are excluded,\nas shown in section IV below.\nLet’s consider the case of a single-mode excitation of\nthe acoustic mode, as it is observed in the simulations\nshown in this paper. In this case, pop= 0andpac=\n¯pac=−γac\nQac>0. Therefore the magnetization oscillates\nat the frequency fof the excited acoustic mode, which\nis given by :\n2πf=ωstt= Ωac=ωac+Nac¯pac (15)\nThis equation is equivalent to the phase equation\nof an STO composed of a single-layer (SL) free layer\nas described in previous work5. The power increases\nwith the applied current, and the frequency decreases\nor increases depending on the sign of Nac. Figure 3\nshows the transition between the two regimes, red-shift\n(frequency decrease with the current) and blue-shift\n(frequency increase) with JRKKY =−5×10−4J/m2, by\nchanging the applied field. The frequency is computed\nfrom the extended NLAO model and compared to\nthe frequency obtained from macrospin simulations,\nboth show a transition between red-shift and blue-shift\nat around−75kA/m. The change of regime with\napplied field in an asymmetric SyF was already observed\nnumerically24and experimentally10.\nAs stated, this transition corresponds to Nacchanging\nsign. The value of the non-linear coefficients of Eq. (14)\nis plotted versus applied field Hxin Figure 4 (a). Nac\nchanges signs at around Hx=−75kA/m, which, indeed,\ncorresponds to the red-shift/blue-shift transition. The\nself-sustained oscillations frequency f=ωstt/(2π)versus9\nFigure 3. Self-sustained oscillations frequency versus applied\ncurrentIwithJRKKY =−5×10−4J/m2and for different\napplied fields, from top to bottom : −40kA/m to −90kA/m.\nThe frequency scale is identical in all the panels, from 0 to\n4 GHz. Solid red line : computed from the extended NLAO\nmodel. Dashed blue line : extracted from LLGS simulations.\nBeyond the spin-flop transition, for Hx=−90kA/m, the\nextended NLAO model is not applicable.\nfieldHxatI= 4mA is reported in Figure 4 (b)\nand compared to the acoustic FMR frequency ωac\nand the frequency obtained from the simulations. We\ndifferentiate four regions, from low to high fields : (i)\nbelow the spin-flop field, at −90kA/m, the extended\nNLAO model is not valid. (ii) for higher fields but below\n−75kA/m, the acoustic mode is excited in the blue-shift\nregime, so ωstt> ωac. The discrepancy between the\nfrequency obtained from the extended NLAO model and\nthe simulation is high, as expected because the model is\nnot valid anymore if ωacis small. (iii) above −75kA/m,\nthe acoustic mode is excited, in the red-shift regime,\nωstt< ωac. The frequency computed from the model\nagrees with the simulations. (iv) above −20kA/m,\nthe applied current is too low to excite a mode, the\noscillator is in sub-critical mode. Notice that in the\nvicinity of the field value at which Nacvanishes, the\noscillator frequency does not change much with the\napplied field, in agreement with the simulations. At this\nfunctioning point, the oscillator frequency is not very\nsensitive neither to the applied field, nor to the applied\ncurrent.\nWe showed that the frequency of the self-sustained os-\ncillationscanbepredictedbytheextendedNLAOmodel,\nin the next section the model will be compared to numer-\nical simulations to define its validity range.\nFigure4. (a)Non-linearand(b)linearfrequencytermsversus\napplied field HxforI= 4mA andJRKKY =−5×10−4J/m2.\n(a) Non-linear coefficients: optical Nop(green), acoustic Nac\n(red),inter-mode T(blue). (b)Linearcoefficients: dottedma-\ngenta line, linear ωac; red solid line, self-sustained oscillations\nfrequency from the model ωstt= Ω ac=ωac+pacNac; dashed\nblue line, self-sustained oscillations frequency from simula-\ntions. The field range is divided in four regions, from low\nto high fields : model non-applicable (NA), blue-shift regime\n(BS), red-shift regime (RS) and no excitation (NE). All values\nare divided by 2πto be in Hz units and not in rad/s.\nC. Single-mode description of the SyF-STO\nTwo sets of simulations are presented, showing the self-\nsustained oscillations frequency versus applied current\nand field for two coupling strengths: (i) Figure 5 in the\nsmallcouplingregime JRKKY =−2×10−4J/m2, (ii)Fig-\nure6the largecouplingregime JRKKY =−5×10−4J/m2.\nIn both figures, the frequency of the m1ycomponent of\nthe magnetization of layer 1 from macrospin simulations\nis plotted in the top panels (a). The frequency computed\nfrom the extended NLAO model is plotted in the bottom\npanels (b). State diagrams for these values of JRKKYare\ndisplayed in reference 25.\nWe observe a qualitative agreement between the model\nand the simulations, especially in the region close to\nthe critical current. First, above the acoustic critical\ncurrentIac\nc(region on the right of the red solid line), the\nmodel predicts self-sustained acoustic-like oscillations,\njust like the simulations (and other publications25).\nJust above the optical critical current Iop\nc(region on\nthe right of the green solid line and on the left of\nthe red solid line), there is no oscillation and the two\nlayersswitch, aspredictedbytheequationsofthepowers.\nThere are also several discrepancies, that will be dis-\ncussed in the following.10\nFigure 5. Frequency of the self-sustained oscillations versus\napplied current and field (a) from macrospin numerical sim-\nulations and (b) from the formulas for the power and phase\nfrom Eq. (15). The RKKY coupling is of −2×10−4J/m2.\nRed (green) solid lines represent Ic(Hx)the vanishing of the\nacoustic (optical) linear dissipative coefficient γac(op). Dotted\nlines correspond to the vanishing of the quadratic dissipative\ncoefficientQac(op).\nFirst, the out-of-plane precession (OPP) region is not\npredicted by the model. OPP are oscillations around\nthe energy maximum, which are not considered in this\nmodel. To describe the OPP, the projection base for the\ncomplexa-coordinates should be changed to the out-of-\nplane axes, instead of the equilibrium in-plane axes, and\nall the coefficients should be computed again.\nSecond, according to the simulations, self-sustained os-\ncillations are expected when the field is larger than the\nspin-flop field. However the extended NLAO model is\nnot valid in the spin-flop region. In fact it is not valid in\nthe vicinity of the spin-flop field either, as it was already\nmentioned. That is why for JRKKY =−2×10−4J/m2,\nFigure 5, the red-shift/blue-shift transition at around\nHx=−45kA/m, is not predicted by the extended\nNLAO model : it is too close to the spin-flop field\nvalue of−50kA/m. On the contrary, for JRKKY =\n−5×10−4J/m2, Figure 6, the red-shift/blue-shift tran-\nsition at around Hx=−70kA/m, with a spin-flop field\nat−90kA/m, is well predicted by the extended NLAO\nmodel.\nLast, the model predicts a much larger region of oscil-\nlations than the simulations. In the region on the right of\nthe optical critical current Iop\nc(green solid line), the dif-\nFigure 6. Same as Figure 5 with an RKKY coupling of −5×\n10−4J/m2.\nference between the model and the simulations becomes\nreally important. This was also shown in Figure 3. In\nthis region, the power is large, which is a known limit\nfor the validity of the NLAO model. But there could be\nanother explanation, because in this region, the model\npredicts a single-mode excitation with pop= 0, whereas\nthe simulations show that popdoes not vanish (not shown\nin the figures).\nTo explain the failure of the model in this region, we\npropose to study the influence of other terms that we\nfirstdiscardedinthemodel, namelythelinearcoefficients\nfrom the dissipative part that are non-resonant. The lin-\near terms are important corrections as they depend lin-\nearly in the powers, contrary to higher order terms. Also\nthey can be easily computed, which is not the case of\nhigher order terms.\nIV. CORRECTION DUE TO NON-RESONANT\nTERMS\nAs was shown in Section IIIC, Eq. 10 cannot capture\nall the features of the dynamics, in particular the fre-\nquency versus current. Therefore, in order to obtain a\nbetter description of the phase, we also include in Eq. 10\nnon-resonant, off-diagonal terms. This leads to the fol-11\nlowing equation:\n˙bop=−(iΩop+ Γop)bop−˜γopb†\nop−˜ϑopbac−ϑopb†\nac\n˙bac=−(iΩac+ Γac)bac−˜γacb†\nac−˜ϑacbop−ϑacb†\nop(16)\nHere Γop=γop+Qoppop+Roppacis the optical dissi-\npative part with only resonant terms from equation (9).\nIdentically, Γac=γac+Qacpac+Racpopis the acoustic\nresonant dissipative part. For the conservative part,\nΩop=ωop+Noppop+TpacandΩac=ωac+Nacpac+Tpop.\nThe coefficients of the non-resonant terms\n(˜γop,˜γac,ϑop,ϑac,˜ϑop,˜ϑac) are independent of the\npowers; for simplicity, we take the coefficients to be real,\nbut taking into account the imaginary part does not\nchange the general conclusions.\nThe equations for the amplitude and phase rewrite as :\n˙pop=−2(Γop+ ˜γopcos(2φop))pop\n−2√poppac/parenleftbig˜ϑopcos(φop−φac) +ϑopcos(φop+φac)/parenrightbig\n˙pac=−2(Γac+ ˜γaccos(2φac))pac\n−2√poppac/parenleftbig˜ϑaccos(φop−φac) +ϑaccos(φop+φac)/parenrightbig\n(17)\n˙φop= Ωop+ ˜γopsin(2φop)\n+/radicalbiggpac\npop/parenleftbig˜ϑopsin(φop−φac) +ϑopsin(φop+φac)/parenrightbig\n˙φac= Ωac+ ˜γacsin(2φac)\n+/radicalbiggpop\npac/parenleftbig˜ϑacsin(φac−φop) +ϑacsin(φop+φac)/parenrightbig\n(18)\nThe equations including the non-resonant terms are\nmore complicated, therefore each term will be treated\nseparately.\nWe first present a qualitative interpretation of each term\nand then evaluate its effect on the dynamics in Figure 7 :\nLLGS equation (Eq. (2)), extended NLAO model with\nonly resonant terms (Eq. (10)), with the addition of the\nlinear dissipative terms (Eq. (16)) and with all the terms.\nA. Inter-mode phase locking\nAn important disagreement between the LLGS sim-\nulation (Eq. (2)) and equation (10) is the phase of the\nnon-excited mode, as can be seen from the comparison\nof Fig. 7 (a) and 7 (c). In section IIIB, it was shown\nthat without the non-resonant terms the two modes have\ndifferent frequencies. However, in the LLGS simulations,\nFig. 7 (a), the two modes are locked, they have the same\nfrequency (although they can have an opposite sign19).\nThis discrepancy can be corrected by including the terms\nwiththe ˜ϑacand˜ϑopcoefficients, asisshowninFig.7(d).\nLet’s suppose that only an acoustic-like mode is ex-\ncited, buttheopticalmodedoesnotvanishtotally( pop≈0andpac= ¯pac). The powers are considered to be con-\nstant.\nThe differential equation for the phases of the two modes\nare :\n˙φop= Ωop+/radicalbiggpac\npop˜ϑopsin(φop−φac)(19)\n˙φac= Ωac+/radicalbiggpop\npac˜ϑacsin(φac−φop)(20)\nIn the acoustic phase equation (20), the second term\nis negligible compared to the constant frequency Ωac\nbecause of the powers ratio, so in the first order, the\nacoustic mode has a constant frequency ˙φac= Ωac, so\nφac= Ωact. However, in the optical phase equation (19),\nthe second term on the right-hand-side is dominant, also\nwith respect to the left-hand-side. This leads to the rela-\ntionsin(φop−φac) =/radicalbiggpop\npac/parenleftbigg˙φop−Ωop\n˜ϑop/parenrightbigg\n≈0, so in the\nfirst order, φop≈Ωact, orφop≈π+ Ωact. This means\nthat the frequency of the non-excited mode is locked to\nthe frequency of the excited mode in the supercritical\nregime.\nAt the second order, the phase difference is given approx-\nimately by :\nφop−φac=/radicalbiggpop\npac/parenleftbiggΩac−Ωop\n˜ϑop/parenrightbigg\n+kπwithk∈Z\n(21)\nSimilarly, the terms with the ϑacandϑopcoefficients\nare responsible for a locking with opposite frequency\n(same absolute frequency, but opposite phase sign), of\nthe form :φop+φac≈0, withφac(t) = Ωact.\nIf both ˜ϑopandϑopare included simultaneously, there\nis a competition between the two terms for the locking\nof the non-excited mode, to the same or the opposite\nfrequency as the excited mode. The resulting relation\nbetween the two phases is more complicated then. How-\never, regarding the time-average of the frequency, the\nnon-excited mode is locked to the frequency of the ex-\ncited mode if|˜ϑop|>|ϑop|, and to the opposite frequency\nif|˜ϑop|<|ϑop|. In other words, the coefficient with the\nhighest value (in norm) determines the type of locking,\ndirect or opposite. An example for opposite frequency\nlockingistheself-polarizedconfigurationdiscussedinref-\nerence 19.\nB. Power oscillations and second harmonics\nSecond, let’s focus on the term with the ˜γaccoefficient\n(itwillbesimilarforthetermin ˜γop). Weconsidera pure\nsingle-mode excitation of the acoustic mode, so pop= 0.\nNote that this analysis is valid for any single-mode non-\nlinear oscillator equation, including the SL case.12\nWithout the other non-resonant terms, the power and\nphase equations of the acoustic mode write :\n˙pac=−2Γacpac−2˜γaccos(2φac)pac\n˙φac= Ωac+ ˜γacsin(2φac)\nIn the assumption that the perturbation due to the ˜γac\nterm is small, one can use Lindstedt’s series to solve this\nsystem of equations26. If/epsilon1=˜γac\nΩacis small, then the\npowerpacand phaseφaccan be written as power series\nof/epsilon1:pac=p0+/epsilon1p1andφac=φ0+/epsilon1φ1. In the zeroth\norder,p0= ¯pacandφ0=¯Ωact, with ¯Ωac=ωac+Nac¯pac.\nIn the first order, the equation for the power deviation\np1and phase deviation φ1are :\n˙p1=−2¯pacQacp1−2¯pac¯Ωaccos(2 ¯Ωact)\n˙φ1=Nacp1+¯Ωacsin(2 ¯Ωact)\nWe use the fact that ¯pacQac=−γac/lessmuch¯Ωac, so the first\nterm on the right-hand side of the power equation is ne-\nglected. Therefore, in the first order and in the perma-\nnent regime, the power pacwrites :\npac(t) = ¯pac/parenleftbigg\n1−˜γac\nΩacsin(2 ¯Ωact)/parenrightbigg\nUp to the first order, the phase φacis given by :\nφac(t) =¯Ωact−˜γacωac\n2¯Ω2accos(2 ¯Ωact)\nTherefore the term ˜γacgives rise to oscillations of the\npower but also a second harmonics in the frequency spec-\ntrum. As a consequence, it also contributes to the STO\nsynchronization by an AC current on the second harmon-\nics. Notice that this term is also present in STO based on\na SL free layer but was omitted in previous descriptions4.\nC. Simulations and trajectories\nThe effect of the non-resonant terms on the dynamics\nis best seen by simulating the different equations.\nOn Fig. 7, we compare the simulations of different\nequations and performed in different coordinate sys-\ntems, and projected afterwards in the (pop,pac,φop,φac)-\ncoordinates for comparison. In Fig. 7 (a), the simulation\nis performed in the (m1x,m1y,m1z,m2x,m2y,m2z)-\ncoordinates, like the usual LLGS simulations, according\nto Eq. (2).\nIn Fig. 7 (b), the simulation is performed in the com-\nplexa-coordinates, using equation (8). The trajectory is\nvery similar to the LLGS trajectory. That is because the\nterms of order superior to 3 in (a1,a2)were dropped after\nthe canonical transformation (m1,m2)−→(a1,a2)and\nwith powers of the order of 10−2, this approximation isperfectly valid.\nIn Fig. 7 (c-e), the simulations are performed in the com-\nplexb-coordinates, from Equation (16). In Fig. 7 (c), all\nthe off-diagonal terms are omitted (which corresponds to\nEq.(10)). Thetrajectoryexhibitsaconstantfiniteacous-\ntic power, a vanishing optical power, and instantaneous\nfrequency for the acoustic and optical mode being con-\nstant but with different values. The constant power and\nfrequency of the acoustic mode are close to the averaged\nvalues computed from the LLGS equation.\nInFig.7(d), ϑop,ϑac,˜ϑopand˜ϑacaretakenintoaccount.\nThe powers are very similar to the powers obtained in\nFig. 7 (c), which justifies the approximation of constant\npowers used in the previous section. The frequency of\nthe non-excited mode, the optical mode, is locked to the\nacoustic frequency. The optical frequency is not constant\nthough, this is because of the competition between the\ntwo types of locking, direct and opposite. But its average\nvalue is close to the value of the acoustic frequency.\nIn Fig. 7 (e), ˜γopand˜γacare also included, so the sim-\nulated equation is exactly Eq. (16). The powers are not\nconstant anymore, but oscillate around the average value\ninstead. Although the average acoustic power is over-\nestimatedcomparedto theLLGSequation (0.043instead\nof 0.036, 20%over-estimated), the average frequencies\nmatch more accurately (-3.40 GHz instead of 3.46 GHz,\n2%under-estimated).\nIn conclusion, Eq. (10) with only resonant terms\npredicts accurately the excitation of the acoustic mode\nfor this set of parameters and it gives a good estimation\nfor the average values of the power and frequency of\nthe excited mode. In order to account for second order\nfeatures, like phase locking of the non-excited mode to\nthe excited mode and first harmonic oscillation, the\ncorrected equation (16) should be used. However, this\ncorrected model is not enough to explain the discrepancy\nwith the LLGS equation in the average power. When\nthe field becomes closer to the spin-flop field, this error\nbecomes so large that extended NLAO model is not\nvalid anymore. As stated in section IIA, the error is\nprobably due to higher order terms but this is out of\nthe scope of this paper. Similarly, the extended NLAO\nmodel fails at large applied currents and this cannot be\nexplained by the correction terms. It is also probably\ndue to higher order terms.\nWith the restrictions of the model of Eq. (10) in mind,\nin the next section we make predictions on how to reduce\nthe generation linewidth of the SyF-STO, which is a very\nimportant parameter for application. The value of the\nlinewidth given by the model were compared to LLGS\nsimulations.13\nFigure 7. Simulations for Hx=−40kA/m,I= 4mA andJRKKY =−5×10−4J/m2performed in the (a) m-variables, (b)\na-variables, (c) b-variables, only with resonant terms from Eq. (10), (d) b-variables with non-resonant terms from Eq. (16) but\n˜γop= ˜γac= 0and (e)b-variables with all non-resonant terms from Eq. (16). The results of the simulations are transformed to\ntheb-coordinates to compare them easily. Insets in (a) and (e) : zoom between 30 and 31 ns. Top panel figures : powers pac\n(red) andpop(green). Bottom panel figures : phase velocity or instantaneous frequency in GHz,∂φac\n∂t(red) and∂φop\n∂t(green).\nV. APPLICATION: REDUCE THE STO\nLINEWIDTH\nA. Thermal noise\nSo far, the system was supposed to be at zero tem-\nperature, however stochastic fluctuations arise at non-\nzero temperature. The effect of these fluctuations can\nbe estimated in regions where the single-mode approx-\nimation is valid. We consider single-mode acoustic-like\nself-sustained oscillations, but the same reasoning apply\nto any single-mode non-linear oscillator.\nWith finite temperature, the power and phase of the os-\ncillator are given by :\n˙pac=−2pac(γac+Qacpac) +/radicalbig\n4pacDacηp(22)\n˙φac=ωac+Nacpac+/radicalBigg\nDac\npacηφ (23)\nWhereηpandηφrepresent white Gaussian noise with\nnormalized variance and the diffusion coefficient Dacis\ndefined by :\nDac= Γ+\nacωT\nΩacwithωT=γ0kBT\n2Mwith Γ+\nacthe positive damping (without the contribution\nfrom the STT) computed at ¯pacandΩac=ωac+Nac¯pac.\nBecause of the thermal noise, the auto-oscillator\nexhibits a finite generation linewidth ∆ω, typical of a\nnon-linear single-mode oscillator6,7. The spectral density\ncan be Lorentzian or Gaussian depending on the value\nof the damping rate of the power fluctuations (or power\nrelaxation rate) Γp= ¯pacQac. The characterization of\na non-linear single-mode oscillator in the presence of\nthermal noise is detailed in Appendix C.\nIf the correlation time of the power fluctuations ( 1/Γp)\nissmallcomparedtothecharacteristicphasedecoherence\ntime(theinverseofthegenerationlinewidthbeingagood\nestimation), ∆ω/lessmuchΓp, the spectral density is Lorentzian\nand the full width at half-maximum (FWHM) ∆ωLis\ngiven by :\n∆ωL= ∆ω0/parenleftbig\n1 +ν2\nac/parenrightbig\n(24)\nwithνac=Nac/Qacand ∆ω0= Γ+\nacωT\n¯pacΩac(25)\nWhere ∆ω0is thelineargeneration linewidth and νacis\nthe normalized non-linear frequency shift coefficient.14\nOn the other hand, if the correlation time of the power\nfluctuations is much larger than the decoherence time,\n∆ω/greatermuchΓp, the spectral density is Gaussian with standard\ndeviation ∆ωGgiven by :\n∆ωG=|νac|/radicalbig\n∆ω0Γp (26)\nThe FWHM is given by√\n8 ln 2 ∆ωG.\nB. Key parameters to the linewidth\nThe expressions of Eq. 24 and 25 identify three pa-\nrameters that can be changed to reduce the value of\nthe linewidth to make functional devices : (i) increase\nthe power relaxation rate Γp, (ii) decrease the linear\nlinewidth ∆ω0and (iii) decrease the normalized non-\nlinear parameter νac.\n•The power relaxation rate Γp= ¯pacQac=|γac|, is\nproportional to the difference between the applied\ncurrent and the critical current Ic. An analytical\nexpression of γacis given in reference 14. In order\nto increase Γpwithout increasing Ic, the absolute\nvalue of the slope of |γac|versusIshould be in-\ncreased without increasing |γac|atI= 0.\n•The linewidth is proportional to the square of the\nnormalized non-linear parameter νac(if the nor-\nmalized non-linear parameter is large, which is the\ncase for STOs). Therefore, one way of reducing the\nlinewidth would be to reduce the non-linear param-\neterNacto zero. For the SyF structure discussed\nhere, this is the case at the transition from the red-\nshift to the blueshift regime. At the transition, the\nlinewidthisequaltothelinearlinewidthvalue ∆ω0.\nIn SL-STO, the vanishing of the non-linear param-\neter can be achieved by changing the equilibrium\nmagnetic state from in-plane along the easy axis to\nin-plane along the hard axis or out-of-plane4. This\nusually requires an external field. In SyF-STO, the\nvanishing of νaccan be achieved by applying an\nin-plane magnetic field along the easy axis. Such\na magnetic field can be generated by the dipolar\nfield from another magnetic layer with the same\neasy axis direction. Notice that a vanishing non-\nlinear parameter Nacmeans that the frequency of\nthe STO becomes independent of its power, and\nthen of the applied current; this loss of tunability\ncan be detrimental for applications. The synchro-\nnization bandwidth with an external signal is also\nproportional to the normalized non-linear parame-\nterνac5, so it should not be too small.\n•The linear linewidth is inversely proportional to\nthe geometrical mean magnetic volume M(see\nEq. (25)). With a SL, the critical current is pro-\nportional to the magnetic volume, so it is counter-\nproductive to increase it. For a SyF however, one\nFigure 8. Power relaxation rate Γpat constant super-\ncriticalityζ= 1andHx= 0versus RKKY coupling energy\nby area, plotted for different layer thicknesses (in nm). The\nother layer properties are the same as in Table I without ap-\nplied field,Hx= 0.Γpis divided by 2πto be expressed in Hz\ninstead of rad/s.\ncan think of a thin layer subjected to the spin-\ntransfer torque from the reference layer, coupled to\na thick layer not subjected to spin transfer torque.\nThus the critical current remains low, whereas the\nmean magnetic volume is increased.\nIn the next sections, we give some ideas about improving\nthese three parameters using a SyF-STO.\nC. Dependence of Γpon the coupling strength\nFirst, we study the variation of the power relaxation\nrate Γpwith some parameters of the SyF. However, be-\ncause Γpis related to the critical current Ic, we need\nto somehow normalize its value. To start, the super-\ncriticalityζis used instead of the current :\nζ=I−Ic\nIc\nUsing this normalized quantity, one can compare the\nvalues of Γpat twice the critical current value, which\ncorresponds to ζ= 1.\nThe applied field dependence of Γpis non-trivial but\nits value at zero field, Hx= 0, is interesting for ap-\nplications. The value of Γpat zero field, for the same\nsuper-criticality ζ= 1, is plotted in Figure 8 for different\nthicknesses of the two layers. It shows that Γpincreases\nwith the RKKY coupling strength, although it remains\nin the same order of magnitude as with a single layer\n(asymptotic value for JRKKY→0).\nD. Vanishing of the non-linear parameter Nac\nBecause of the quadratic dependence of the linewidth\non the normalized non-linear parameter νac, the most\neffective action to reduce the oscillator linewidth is to15\nFigure 9. Linewidth of m1y(yellow diamonds) from LLGS\nsimulations at 300 K, compared to the linewidth (solid red\nline), linearlinewidth (dotted red line) and Γp(dashed blue\nline) computed from the extended NLAO model, versus ap-\nplied field for a current of I= 4mA andJRKKY =−5×\n10−4J/m2.\ndecreaseNacby applying an in-plane field so the oscil-\nlator is excited close to the transition between red-shift\nand blue-shift.\nFigure 9 shows a comparison of the linewidth from LLGS\nsimulations at 300 K and from the extended NLAO\nmodel. The linewidth is plotted versus applied field,\natI= 4mA andJRKKY =−5×10−4J/m2. For\nthe simulations, the linewidth is calculated from a fit\nto a Lorentzian function. We observe a decrease of the\nlinewidth of almost two orders of magnitude between\nHx= 0andHx=−70kA/m. The linewidth decrease is\nassociated to the vanishing of the non-linear parameter\nNac. For small fields, |Hx|<50kA/m, the linewidth is\nmuch larger than the power relaxation rate, which cor-\nresponds to a Gaussian spectrum. On the other hand,\naroundHx=−70kA/m, the spectrum has a Lorentzian\nprofile. In the simulations, the spectrum appears to be\nindeed Lorentzian around Hx=−70kA/m. It is difficult\nto conclude about the line shape at lower absolute field\nvalue, though, because the noise is too large and both\nprofiles interpolate well the simulated spectrum.\nFigure 10 shows the linewidth versus field for a low cou-\npling,JRKKY =−2×10−4J/m2, and forI= 3mA.\nAs was shown above, the model does not predict a van-\nishing ofNac, therefore the predicted linewidth remains\nlarge in the whole field range. However, the macrospin\nsimulations show a redshift/blueshift transition at Hx=\n−45kA/m and a decrease of the linewidth to its linear\nvalueatthisfield. Infact, atthisfield, thefrequencydoes\nnot change with the applied current. In a single-mode\nmodel, it means that the phase does not depend on the\npower, so the linewidth is given by the linear linewidth\nalone. Therefore, in the low coupling regime, the os-\ncillation looks like it is single-mode, according to the\nmacrospin simulations, but the extended NLAO model\nis not sufficient to estimate the characteristic parameters\nof the oscillator.\nFigure 10. Same as Figure 9 with JRKKY =−2×10−4J/m2\nandI= 3mA.\nFigure11. Linewidthof m1yversusfield, comparisonbetween\na 2 nm thick single layer (blue) and a 2 nm layer coupled with\na 20 nm thick layer (red-orange), separated by (a) 1 nm and\n(b) 20 nm spacer. Symbols : LLGS simulations, solid lines :\nextended NLAO model, dotted lines : linearlinewidth from\nthe extended NLAO model.\nE. Coupling to a thick layer\nFinally, the last parameter that can be tuned to\nreduce the linewidth is the linear linewidth ∆ω0. The\nlinear linewidth does not depend much on the coupling\nstrength, but more on the magnetic volume, as stated\nbefore. In order to increase the total magnetic volume\nand keep a reasonable critical current, we can imagine a\nthinlayerof2nmcoupledtoathicklayerof20nm. With\nthis geometry, where the layers are very asymmetric,\nthe coupling strength plays an important role. Contrary16\nto the asymmetric case studied previously where the\nnon-linear parameter vanishes with the combination\nof asymmetric layers and strong coupling25, in the\nfollowing example, the non-linear parameter is reduced,\nso the linewidth is decreased, although not to the level of\nthelinearlinewidth. However, the linewidth reduction\nhappens at lower fields, more suitable for application,\nthan in the previous case.\nThe SyF of this example is compared to a nano-pillar\nbased on a single free layer. The SL-STO is composed\nof three layer : (1) a reference layer with in-plane\nfixed magnetization, a spin polarization of 0.3and\ncompensated dipolar fields (the total stray field is zero),\n(2) a tunnel barrier, and (3) a 2 nm thick free layer, with\nsaturation magnetization of 1×106A/m and damping\nconstant of 0.02. The nano-pillar has an elongated shape\nof150×100nm, giving a shape anisotropy to the free\nlayer along the x-axis. The SyF-STO of study comprises\nthe same SL nano-pillar, plus two additional layers : (4)\na spacer of variable thickness, 1 nm or 20 nm, and (5)\na 20 nm thick free layer with saturation magnetization\nof1×106A/m and damping constant of 0.02. The\nmagnetizations of the thick and the thin layers are\ncoupled through dipolar field, whose strength is lower or\nhigher depending on the thickness of the spacer. For the\ntwo cases, strong and weak coupling, the coefficients of\nequation (1) take the values :\nStrong coupling ( tMgO= 1nm) :\n(˜Dx,˜Dy,˜Dz)/S= (−1.9,−2.9,4.8)×10−4J/m2\nWeak coupling ( tMgO= 20nm) :\n(˜Dx,˜Dy,˜Dz)/S= (−0.9,−1.4,2.3)×10−4J/m2\nDue to the shape anisotropy, the magnetization of the\nthick layer is more stable than that of the thin layer, but\nit is still free to move. Like for the other stacks stud-\nied in this paper, the current is spin polarized between\nthe reference layer and the 2 nm thin layer, but it is\nconsidered unpolarized at any other point, including be-\ntween the thin and thick layers. The simulations were\nperformed at 300 K and the linewidth is computed do-\ning a Lorentzian fit of the power spectral density of m1y,\nthe magnetization of the thin layer along the y-axis. The\nlinewidth computed from the extended NLAO model and\nextracted from the simulation are showed in Figure 11.\nWhen the thin and thick layers are separated by 20 nm\nso they are weakly coupled, Figure 11 (a), the linewidth\nis of the same order of magnitude with or without the\nthick layer, in the hundreds of MHz range. Around\n10kA/m, which is the coupling field (at which the thick\nand thin layers have the same FMR frequency), the ex-\ntendedNLAOmodelpredictsanincreaseofthelinewidth\nabove the SL value, that is not observed in the simula-\ntions. On the contrary, the simulations show a decreaseof the linewidth around 10kA/m that we cannot explain.\nOverall, the value of the SyF linewidth is essentially com-\nparable to the value of the SL linewidth.\nIn the strongly coupled case, with a 1 nm spacer, Fig-\nure 11 (b), below the coupling field (around -10 kA/m),\nthe model predicts a reduction of the linewidth of one or-\nderofmagnitudebetweentheSyFandtheSLcase; above\nthe coupling field, an increase of the linewidth is pre-\ndicted. The simulations show a decrease of the linewidth\nofalmostoneorderofmagnitudefortheSyFcomparedto\nthe SL case for fields smaller than the coupling field, with\na minimum of 5 MHz at -25 kA/m, in agreement with\nthe model. Notice that the decreased linewidth is still\none order of magnitude larger than the linearlinewidth.\nAround the coupling field, the SyF and SL linewidths are\nequivalent, around 100 MHz. Above the coupling field,\nthe linewidth of the SyF is half the linewidth of the SL,\nin disagreement with the model.\nIn conclusion, we observe a reduction of the linewidth\nwhen a thin layer is strongly coupled to a thick layer.\nThe linewidth reduction occurs for all the fields except\nfor the coupling field, at which the linewidth value is as\nhigh as for a single layer.\nVI. CONCLUSION\nWe presented an extension of the NLAO model to de-\nscribe the self-sustained oscillations of a SyF composed\nof two layers coupled with RKKY coupling, dipolar cou-\npling and mutual STT. The analysis was restricted to\nthe plateau region of the SyF, where the two layers are\naligned along the same direction at equilibrium, paral-\nlel or anti-parallel. However, nothing prevents one from\napplying the same analysis to arbitrary initial configura-\ntions (and an arbitrary number of layers), by taking into\naccount a transverse field for instance, although the di-\nagonalization of the hamiltonian matrix would be more\ncomplicated and only numerical solution would be avail-\nable.\nIn the extended model, the SyF dynamics is described\nby two coupled complex non-linear equations, which cor-\nrespond, in the linear regime, to the acoustic and optical\nmode. Inthispaper, wefocusedonSyFswithfixedexter-\nnal polarizer and, for the set of parameters that we chose,\nonly one mode is excited at a time, the acoustic-like self-\noscillation. Therefore, the dynamics can be described by\na single mode power and a phase equation, as in the case\nof a single layer. It means that the self-sustained oscil-\nlations are defined by a constant power, resulting from\nthe balance between natural damping and STT. The fre-\nquency consists of a linear part and a non-linear part,\nproportionaltothepowerandtothenon-linearfrequency\nshiftNac. Identically, the linewidth of the power spectral\ndensity consists of a linear part and a non-linear part.\nIt was found that with a strong coupling and if the\ntwo layers are asymmetric, for instance if they have dif-\nferent thicknesses, the non-linear frequency shift Naccan17\nbe reduced strongly, so the linewidth is also strongly re-\nduced of one order of magnitude. In particular cases, Nac\ncan even vanish at a given field, which corresponds to a\ntransition between a red-shift and a blue-shift frequency\nversus current dependency. At this field, the linewidth is\nreduced to its linearlinewidth value, which is a reduction\nof almost two orders of magnitude. The power relaxation\nrateΓpwas not found to change much compared to the\nvalues found for a single layer STO.\nThis work confirms the robustness of the NLAO model\ntodescribesmalloscillationsofthemagnetizationaround\nthe equilibrium and it shows that it can be extended to\nseverallayers. Italsopresentedarelativelysimplesystem\nto study the interaction between oscillating modes and\nwe hope it can be extended to more general cases.\nACKNOWLEDGMENTS\nThisworkwassupportedbytheEuropeanCommission\nundertheFP7programNo.316657SpinIcurandtheFP7\nprogram No. 317950 MOSAIC.18\nAppendix A: Hamiltonian diagonalization :\ntransformation a-b\nThe expression of the coefficients of the transformation\nmatrixTabare given by the 6 angles : φj,ψjandθjfor\ni= (op,ac).\nFirst the angles φj(forj= (op,ac)) are computed :\nR+\nj= (A1+ωj)(A2+ωj)−D2\n12\nusj=A1−ωj+D12+C12D12−B1(A2+ωj)\nR+\nj(B1+C12)\n+B1D12−C12(A1+ωj)\nR+\nj(B2+C12)\nucj=A2−ωj+D12+B2D12−C12(A2+ωj)\nR+\nj(B1+C12)\n+C12D12−B2(A1+ωj)\nR+\nj(B2+C12)\nunj=/radicalBig\nus2\nj+ uc2\nj\nsinφj=usj\nunjcosφj=ucj\nunj\nNext the angles ψj(forj= (op,ac)) :\nR−\nj= (A1−ωj)(A2−ωj)−D2\n12\nvsj=A1+ωj+D12+C12D12−B1(A2−ωj)\nR−\nj(B1+C12)\n+B1D12−C12(A1−ωj)\nR−\nj(B2+C12)\nvcj=A2+ωj+D12+B2D12−C12(A2−ωj)\nR−\nj(B1+C12)\n+C12D12−B2(A1−ωj)\nR−\nj(B2+C12)\nvnj=/radicalBig\nvs2\nj+ vc2\nj\nsinψj=vsj\nvnjcosψj=vcj\nvnj\nAnd finally, the angles θj(forj= (op,ac)) are com-\nputed :\nF1=A1−B1−C12F2=A2−B2−C12tanhθj=−cosφj(F1−ωj)−sinφj(F2−ωj)\ncosψj(F1+ωj)−sinψj(F2+ωj)\nAppendix B: Coefficients of the dissipative part\nThe dissipative part is expressed as a power series in\nthea-coordinates, truncated after the cubic term :\nFai=/summationdisplay\np,q,r,sfp,q,r,s\naia1pa2qa†\n1ra†\n2sfori= 1,2\nWe use the following notations :\nν1=−mγ0\n2M~\n2|e|Iη1ν2=−mnγ0\n2M~\n2|e|Iη2\nν21= +γ0\n2M~\n2|e|Iη21ν12=−γ0\n2M~\n2|e|Iη12\nκ1=−mγ0\n2M~\n2|e|Iβ1κ2=−mnγ0\n2M~\n2|e|Iβ2\nκ21= +γ0\n2M~\n2|e|Iβ21κ12=−γ0\n2M~\n2|e|Iβ12\nHence the non-vanishing coefficients of Fa1andFa2with\nindices (p,q,r,s )are given by ( iis the imaginary unit,\ni2=−1) :\nFa1:\n(1,0,0,0) :α1A1+ 2nβν 21+ 2βν1−2inβκ 21−2iβκ1\n(0,1,0,0) :α1D12−(1 +n)ν21+i(1 +n)κ21\n(0,0,1,0) :α1B1\n(0,0,0,1) :α1C12+ (1−n)ν21−i(1−n)κ21\n(2,0,1,0) :−α1βA1+ 2α1U1−2nβ2ν21−2β2ν1\n(1,1,0,1) :α1W12−4nν21+ 4inκ21\n(0,2,0,1) :α1Z21+1 +n\n2βν21−i1 +n\n2βκ21\n(0,1,0,2) :α1Y21−1−n\n2βν21+i1−n\n2βκ21\n(1,1,1,0) : 2α1Z12+ (1 +n)βν21−iβ(1 +n)κ21\n(1,0,1,1) : 2α1Y12−(1−n)βν21+iβ(1−n)κ21\n(2,0,0,1) : 3α1Z12+ 31 +n\n2βν21−i1 +n\n2βκ21\n(2,1,0,0) : 3α1Y12−31−n\n2βν21+i1−n\n2βκ21\n(1,0,2,0) : 3α1V1\n(3,0,0,0) : 3α1V119\nFa2:\n(1,0,0,0) :α2D12−(1 +n)ν12+i(1 +n)κ12\n(0,1,0,0) :α2A2+2n\nβν12+2\nβν2−i2n\nβκ12−i2\nβκ2\n(0,0,1,0) :α2C12+ (1−n)ν12−i(1−n)κ12\n(0,0,0,1) :α2B2\n(0,2,0,1) :−α2\nβA2+ 2α2U2−2n\nβ2ν12−2\nβ2ν2\n(1,1,1,0) :α2W12−4nν12+ 4inκ12\n(2,0,1,0) :α2Z12+1 +n\n2βν12−i1 +n\n2βκ12\n(1,0,2,0) :α2Y12−1−n\n2βν12+i1−n\n2βκ12\n(1,1,0,1) : 2α2Z21+1 +n\nβν12−i1 +n\nβκ12\n(0,1,1,1) : 2α2Y21−1−n\nβν12+i1−n\nβκ12\n(0,2,1,0) : 3α2Z21+ 31 +n\n2βν12−i1 +n\n2βκ12\n(1,2,0,0) : 3α2Y21−31−n\n2βν12+i1−n\n2βκ12\n(0,3,0,0) : 3α2V2\n(0,1,0,2) : 3α2V2\nAppendix C: Thermal noise and Fokker-Planck\nequation\nThermal noise is introduced in Eq. (10) in the form :\n˙bop+bop(iΩop+ Γop) =/radicalbig\n2Dopηop\n˙bac+bac(iΩac+ Γac) =/radicalbig\n2Dacηac(C1)\nThe noise amplitudes DopandDac, also called diffusion\ncoefficients, are not constant and depend on the mode\npowers :Dop(bop,bac)andDac(bop,bac), but this depen-\ndence is omitted for clarity. They will be determined\nlater. Ωop,Ωac,ΓopandΓacare the conservative (for op-\ntical and acoustic modes) and the dissipative determin-\nistic coefficients. They also depend on the mode powers.\nηopandηacare two independent white noise sources with\nzero mean and correlators given by :\n/angbracketleftηi(t)/angbracketright= 0, for i∈(op, ac)\n/angbracketleftηi(t)ηj(t/prime)/angbracketright= 0, for i,j∈(op, ac)2\n/angbracketleftηi(t)¯ηj(t/prime)/angbracketright=δijδ(t−t/prime), fori,j∈(op, ac)2\nThe expressions of the diffusion coefficients are de-\ntermined by insuring that the equilibrium probability\ndensity function (PDF) for the powers and phase re-\nduces to the Boltzmann distribution without applied cur-\nrent5. Considering the Stratonovich stochastic differen-\ntial equation (SDE) (C1), the time evolution of the PDFP(pop,pac,φop,φac,t)is given by the following Fokker-\nPlanck (FP) equation :\n∂P\n∂t−∂\n∂pop(2popΓopP)−∂\n∂pac(2pacΓacP)\n+∂\n∂φop(ΩopP) +∂\n∂φac(ΩacP)\n=∂\n∂pop/parenleftbigg\n2popDop∂P\n∂pop/parenrightbigg\n+∂\n∂pop/parenleftbigg\nP∂\n∂pop(popDop)/parenrightbigg\n+∂\n∂pac/parenleftbigg\n2pacDac∂P\n∂pac/parenrightbigg\n+∂\n∂pac/parenleftbigg\nP∂\n∂pac(pacDac)/parenrightbigg\n+Dop\n2pop∂2P\n∂φ2op+Dac\n2pac∂2P\n∂φ2ac\nHere, we considered that the diffusion coefficients depend\nonlyonthemodepowers. Thetermsintheleft-hand-side\ncome from the deterministic equation, or drift, whereas\nthe terms in the right-hand-side represent the thermal\ndiffusion. At equilibrium/parenleftbigg∂P\n∂t= 0/parenrightbigg\n, the PDFP0is a\nuniform distribution for the phases, so we can remove the\nlast two drift terms of the left-hand-side. Moreover, the\nsecond and fourth diffusion terms of the right-hand side\nshould be compensated by two terms of drift that are\nusually neglected. They arise from the renormalization\nof the multiplicative noise terms27(see reference28where\nthese extra drift terms are included for a SL free layer).\nThe extra drift terms can be incorporated in (C1) to give\nthe correct equation in Stratonovich form :\n˙bop+bop(iΩop+ Γop) +fopbop=/radicalbig\n2Dopηop\n˙bac+bac(iΩac+ Γac) +facbac=/radicalbig\n2Dacηac\n(C2)\nWith :\nfop=−1\n2pop∂(popDop)\n∂pop\nfac=−1\n2pac∂(pacDac)\n∂pac\nInterestingly, these extra drift terms contribute only to\nthe power equations. In particular, they are responsible\nfor the non-zero average power below threshold (when\nsolving ˙p= 0,p= 0is not a solution anymore).\nAfter eliminating the extra drift terms, the FP equa-\ntion at equilibrium reduces to :\n0 =∂\n∂pop/parenleftbigg\n2popΓ+\nopP0+ 2popDop∂P0\n∂pop/parenrightbigg\n+∂\n∂pac/parenleftbigg\n2pacΓ+\nacP0+ 2pacDac∂P0\n∂pac/parenrightbigg\nWhere Γ+\nopandΓ+\nacare the dissipative terms at zero ap-\nplied current, i.e. the natural damping.20\nA solutionP0(pop,pac)of the former equation is :\nP0=Z−1exp/parenleftbigg\n−/integraldisplaypop\n0Γ+\nop\nDopdpop−/integraldisplaypac\n0Γ+\nac\nDacdpac/parenrightbigg\nWhereZis a normalization constant. The equilibrium\nPDF should correspond to the Boltzmann distribution,\nwhichisequalto Z/prime−1exp/parenleftbigg\n−E\nkBT/parenrightbigg\n, whereZ/primeisanother\nnormalization constant, Eis the energy of the system as\ndefined in Eq. (1) and Tis the temperature. Then the\ndiffusion coefficients are given by :\nDop= Γ+\nopkBT/parenleftbigg∂E\n∂pop/parenrightbigg−1\n= Γ+\nopωT\nΩop(C3)\nDac= Γ+\nackBT/parenleftbigg∂E\n∂pac/parenrightbigg−1\n= Γ+\nacωT\nΩac(C4)\nWe now consider the self-oscillation regime with a\nsingle-mode excitation of the acoustic mode, with ther-\nmal noise. The stochastic differential equation of the\npower and phase is expressed in the It¯ o form, which is\npreferred when solving analytically stochastic equations\nbecause the solutions are martingales. For clarity, the ac\nindex is dropped on the power pand phaseφ:\n˙p=−2p/parenleftBig\nγac+Qacp+˜fac(p)/parenrightBig\n+/radicalbig\n4pDacηp(C5)\n˙φ=ωac+Nacp+/radicalBigg\nDac\npηφ (C6)\nWhereηp=Re(√\n2ηaceiφac)andηφ=Im(√\n2ηaceiφac)\nare real stochastic variables with zero average and\n/angbracketleftηi(t)ηj(t/prime)/angbracketright=δijδ(t−t/prime), fori,j∈(p,φ).˜fac= 2fac\nis the extra drift term in the It¯ o form, computed from its\nStratonovich form and the diffusion coefficients.\nDue to the extra ˜facterm, the stationary power is dif-\nferent from the power p0without temperature. How-\never, above the threshold, we suppose that the stationary\npower ˜p0is close to the zero-temperature value:\n˜p0=p0(1 +δp0) withδp0/lessmuch1\nIt can be shown that δp0is given by:\nδp0=−˜fac(p0)\np0Qac=∆ω0\nΓp−ν∆ω0\nΩ0Γ+(p∞)\nΓ+(p0)Where ∆ω0=Dac(p0)\np0is thelineargeneration\nlinewidth, Γp=p0Qacis the power relaxation\nrate, Ω0=ωac+p0Nacis the stationary frequency,\nν=Nac/Qacis the normalized non-linear frequency\nshift coefficient and p∞=−ωac\nNac, with positive or\nnegative value. If Nacis negative, p∞corresponds to the\nmaximum oscillation power, for which ωac+Nacp∞= 0.\nAs long as ∆ω0/lessmuchΓpand the oscillation frequency Ω0is\nhigh enough ( Ω0/greatermuchν∆ω0), the effect of the extra drift\nterm can be neglected and ˜p0≈p0.\nThen, weconsiderfluctuationsofthepoweraroundthe\nequilibrium power p0and of the phase around φ0(t) =\nΩ0t:δp=p−p0(withδp/lessmuchp0) andδφ=φ−φ0:\n˙δp=−2p0˜Qδp+/radicalbig\n4p0Dacηp (C7)\n˙δφ=Nacδp+/radicalBigg\nDac\np0ηφ (C8)\nWhere the effective non-linear relaxation rate coefficient\nis˜Q=Qac+∂˜fac\n∂pac/vextendsingle/vextendsingle/vextendsingle/vextendsingle\np=p0.\nThe correction due to the temperature-dependent term\non the non-linear relaxation rate writes as :\n˜Q\nQac−1 =∆ω0\nΓp+ν∆ω0/parenleftbigg2ωac−Ω0\nΩ2\n0/parenrightbiggΓ+(p∞)\nΓ+(p0)\nThe same conditions that assured that δp0/lessmuch1lead to\n˜Q≈Qac.\nBecause the stochastic equations are linear, the power\nand phase fluctuations are Gaussian processes with zero\nmean. There are contributions to the linewidth from the\nphase noise ( ηφ) and from the amplitude noise ( Nacδp).\nNotethattheothermode, theopticalmode, isconsidered\nto be subcritical, so its power is almost zero, and in any\ncase much smaller than the power of the acoustic mode.\nTherefore its contribution to the power spectral density\nis neglected.\nThe power is a weakly stationary process but the phase\nis a non-stationary Gaussian random walk. We obtain\nthe expression of the power variance ∆p2=/angbracketleftδp2/angbracketrightand\nthe phase variance ∆φ2=/angbracketleftδφ2/angbracketright7:\n∆p2=p2\n0∆ω0\nΓp\n∆φ2= ∆ω0/bracketleftbigg\n(1 +ν2)|t|−ν21−e−2Γp|t|\n2Γp/bracketrightbigg\n∗M. Romera is now working at Unité Mixte de Physique,\nCNRS, Thales, Univ. Paris-Sud, Université Paris-Saclay,\n91767 Palaiseau, France1S. I. Kiselev, J. C. Sankey, I. N. Krivorotov, N. C. Emley,\nR. J. Schoelkopf, R. A. Buhrman, and D. C. Ralph, Nature\n425, 380 (2003), ISSN 1476-4687, URL http://www.ncbi.21\nnlm.nih.gov/pubmed/14508483 .\n2W. H. Rippard, M. R. Pufall, S. Kaka, S. E. Russek, and\nT.J.Silva, Physicalreviewletters 92, 027201(2004), ISSN\n0031-9007.\n3M. Tsoi, J. Z. Sun, and S. S. P. Parkin, Physical Review\nLetters93, 036602 (2004), ISSN 00319007.\n4A. Slavin and V. Tiberkevich, IEEE Transactions on\nMagnetics 44, 1916 (2008), ISSN 0018-9464, URL\nhttp://ieeexplore.ieee.org/lpdocs/epic03/wrapper.\nhtm?arnumber=4544933 .\n5A. Slavin and V. Tiberkevich, Magnetics, IEEE\nTransactions on 45, 1875 (2009), ISSN 0018-9464,\nURL http://ieeexplore.ieee.org/xpls/abs{_}all.\njsp?arnumber=4802339http://ieeexplore.ieee.org/\nlpdocs/epic03/wrapper.htm?arnumber=4802339 .\n6J.-V. Kim, V. Tiberkevich, and A. Slavin, Physical Re-\nview Letters 100, 1 (2008), ISSN 0031-9007, URL http:\n//link.aps.org/doi/10.1103/PhysRevLett.100.017207 .\n7V. Tiberkevich, a. Slavin, and J.-V. Kim, Physical Review\nB78, 092401 (2008), ISSN 1098-0121, URL http://link.\naps.org/doi/10.1103/PhysRevB.78.092401 .\n8J. F. Sierra, M. Quinsat, F. Garcia-Sanchez, U. Ebels,\nI. Joumard, A. S. Jenkins, B. Dieny, M. C. Cyrille,\nA. Zeltser, and J. A. Katine, Applied Physics Letters 101\n(2012), ISSN 00036951.\n9B. Georges, J. Grollier, M. Darques, V. Cros, C. Deran-\nlot, B. Marcilhac, G. Faini, and A. Fert, Physical Review\nLetters101, 017201 (2008), ISSN 0031-9007, URL http:\n//link.aps.org/doi/10.1103/PhysRevLett.101.017201 .\n10D. Houssameddine, J. F. Sierra, D. Gusakova, B. Delaet,\nU. Ebels, L. D. Buda-Prejbeanu, M.-C. Cyrille, B. Di-\neny, B. Ocker, J. Langer, et al., Applied Physics\nLetters96, 072511 (2010), ISSN 00036951, URL\nhttp://scitation.aip.org/content/aip/journal/\napl/96/7/10.1063/1.3314282 .\n11T. Nagasawa, K. Kudo, H. Suto, K. Mizushima, and\nR. Sato, Applied Physics Letters 105, 182406 (2014), ISSN\n0003-6951, URL http://scitation.aip.org/content/\naip/journal/apl/105/18/10.1063/1.4901077 .\n12T. Seki, H. Tomita, M. Shiraishi, T. Shinjo, and Y. Suzuki,\nApplied Physics Express 3, 033001 (2010), ISSN 1882-\n0778, URL http://apex.ipap.jp/link?APEX/3/033001/ .\n13A. S. Jenkins, B. Lacoste, G. Geranton, D. Gusakova,\nB. Dieny, U. Ebels, and L. D. Buda-Prejbeanu, Jour-\nnal of Applied Physics 115, 083911 (2014), ISSN\n0021-8979, URL http://scitation.aip.org/content/\naip/journal/jap/115/8/10.1063/1.4866871http:\n//aip.scitation.org/doi/10.1063/1.4866871 .\n14B. Lacoste, L. D. Buda-Prejbeanu, U. Ebels, and B. Dieny,\nPhysical Review B 89, 064408 (2014), ISSN 1098-0121,\nURL http://link.aps.org/doi/10.1103/PhysRevB.89.\n064408.\n15M. Ichimura, T. Hamada, H. Imamura, S. Taka-\nhashi, and S. Maekawa, Journal of Applied\nPhysics109, 07C906 (2011), ISSN 00218979, URL\nhttp://link.aip.org/link/JAPIAU/v109/i7/p07C906/\ns1{&}Agg=doihttp://scitation.aip.org/content/aip/\njournal/jap/109/7/10.1063/1.3549437 .\n16J. L. van Hemmen, Zeitschrift für Physik B CondensedMatter38, 271 (1980), ISSN 0722-3277, URL http://\nlink.springer.com/10.1007/BF01315667 .\n17T. Devolder and K. Ito, Journal of Applied\nPhysics111, 123914 (2012), ISSN 00218979, URL\nhttp://scitation.aip.org/content/aip/journal/jap/\n111/12/10.1063/1.4729776 .\n18V. S. L’vov, Wave Turbulence Under Parametric\nExcitation , vol. 2 of Springer Series in Nonlin-\near Dynamics (Springer Berlin Heidelberg, Berlin,\nHeidelberg, 1994), ISBN 978-3-642-75297-1, URL\nhttp://www.springerlink.com/index/10.1007/\n978-3-642-75295-7http://link.springer.com/10.\n1007/978-3-642-75295-7 .\n19M. Romera, B. Lacoste, U. Ebels, and L. D. Buda-\nPrejbeanu, Physical Review B 94, 094432 (2016),\nISSN 2469-9950, URL http://link.aps.org/doi/10.\n1103/PhysRevB.94.094432 .\n20F. M. de Aguiar, A. Azevedo, and S. M. Rezende, Physical\nReview B 75, 132404 (2007), ISSN 1098-0121, URL http:\n//link.aps.org/doi/10.1103/PhysRevB.75.132404 .\n21J. Hofbauer and K. Sigmund, The theory of evolution\nand dynamical systems. Mathematical aspects of se-\nlection. (CAMBRIDGE UNIVERSITY PRESS, NEW\nYORK, NY, 1988), URL http://scholar.google.\ncom/scholar?hl=en{&}btnG=Search{&}q=intitle:\nThe+Theory+of+Evolution+and+Dynamical+Systems+-+\nMathematical+Aspects+of+Selection{#}0 .\n22I. M. Bomze, Biological Cybernetics 48, 201 (1983),\nISSN 0340-1200, URL http://link.springer.com/10.\n1007/BF00318088 .\n23I. M. Bomze, Biological Cybernetics 72, 447 (1995),\nISSN 0340-1200, URL http://link.springer.com/10.\n1007/BF00201420 .\n24D. Gusakova, D. Houssameddine, U. Ebels, B. Dieny,\nL. Buda-Prejbeanu, M. Cyrille, and B. Delaët, Physi-\ncal Review B 79, 1 (2009), ISSN 1098-0121, URL http:\n//link.aps.org/doi/10.1103/PhysRevB.79.104406 .\n25E. Monteblanco, F. Garcia-Sanchez, D. Gusakova, L. D.\nBuda-Prejbeanu, and U. Ebels, Journal of Applied Physics\n121, 013903 (2017), ISSN 0021-8979, URL http://aip.\nscitation.org/doi/10.1063/1.4973525 .\n26B. Lacoste, L. D. Buda-Prejbeanu, U. Ebels, and B. Dieny,\nPhysical Review B 88, 054425 (2013), ISSN 1098-0121,\nURL http://link.aps.org/doi/10.1103/PhysRevB.88.\n054425.\n27Note1, the diffusion coefficients can be renormalized if the\nSDE is expressed in the It¯ o form, which differs from the\nStratonovich form by an extra drift term. The starting\npoint is a stochastic LLGS equation in the Stratonovich\nform, which is used to describe physical noise. In order to\nsimplify the expression of the noise diffusion terms, one\nmust convert the SDE to the It¯ o form, then simplify the\ndiffusion coefficients, and convert the SDE back to the\nStratonovich form. This process adds two extra drift terms\nthat do not balance each other.\n28T. Taniguchi, Applied Physics Express 7, 053004\n(2014), ISSN 1882-0778, URL http://stacks.\niop.org/1882-0786/7/i=5/a=053004?key=crossref.\n0d01d0a8539d0d8a726bd42873186e86 ." }, { "title": "2401.13617v1.Current_Driven_Domain_Wall_Motion_in_Curved_Ferrimagnetic_Strips_Above_and_Below_the_Angular_Momentum_Compensation.pdf", "content": "Current-Driven Domain Wall Motion in Curved Ferrimagnetic Strips Above and\nBelow the Angular Momentum Compensation\nD. Osuna Ruiz *1,∗O. Alejos2, V. Raposo1, and E. Mart´ ınez1\n1Department of Applied Physics, University of Salamanca, Salamanca 37008, Spain and\n2Department of Electricity and Electronics, University of Valladolid, Valladolid, Spain.\n(Dated: January 25, 2024)\nCurrent driven domain wall motion in curved Heavy Metal/Ferrimagnetic/Oxide multilayer strips\nis investigated using systematic micromagnetic simulations which account for spin-orbit coupling\nphenomena. Domain wall velocity and characteristic relaxation times are studied as functions of\nthe geometry, curvature and width of the strip, at and out of the angular momentum compensation.\nResults show that domain walls can propagate faster and without a significant distortion in such\nstrips in contrast to their ferromagnetic counterparts. Using an artificial system based on a straight\nstrip with an equivalent current density distribution, we can discern its influence on the wall terminal\nvelocity, as part of a more general geometrical influence due to the curved shape. Curved and narrow\nferrimagnetic strips are promising candidates for designing high speed and fast response spintronic\ncircuitry based on current-driven domain wall motion.\nI. INTRODUCTION\nA magnetic domain wall (DW) is the transition re-\ngion that separates two uniformly magnetized domains\n[1]. These magnetic configurations are interesting due to\nfundamental physics, but also due to potential technolog-\nical applications [2, 3]. In fact, during the last decades\nDWs have been at the core of theoretical and experimen-\ntal studies which have provided with a deep understand-\ning of different spin-orbit coupling phenomena [4–8]. For\ninstance, straight stacks where an ultra-thin ferromag-\nnetic (FM) layer is sandwiched between a heavy metal\n(HM) and an oxide (Ox), present perpendicular mag-\nnetic anisotropy (PMA) and therefore, the domains are\nmagnetized along the out-of-plane direction of the stacks:\nup(+⃗ uz) or down (−⃗ uz). DWs in these HM/FM/Ox\nstacks adopt an homochiral configuration due to the\nDzyaloshinskii-Moriya interaction (DMI) [5, 7, 9]. Adja-\ncent DWs have internal magnetic moments along the lon-\ngitudinal direction ( ⃗ mDW=±⃗ ux), and the sense is im-\nposed by the sign of the DMI, which in turns depends on\nthe HM [5]. For left-handed stacks such as Pt/Co/AlO,\nup-down (UD) and down-up (DU) DWs have internal mo-\nments with ⃗ mDW=−⃗ uxand⃗ mDW= +⃗ ux, respectively\n[5]. These DWs are driven with high efficiency by inject-\ning electrical currents along the longitudinal direction of\nthe HM/FM/Ox stack [4]. Due to the spin-Hall effect [5],\nthe electrical current in the HM generates a spin polar-\nized current which exerts spin-orbit torques (SOTs) on\nthe magnetization of the FM layer, and drives series of\nhomochiral DWs which are displaced along the longitudi-\nnal direction ( x-axis). DW velocities of VDW∼500 m/s\nhave been reported upon injection of current densities of\nJHM∼1 TA/m2in Pt/Co/AlO [4]. Consequently, these\nstacks have been proposed to develop highly-packed mag-\nnetic recording devices, where the information coded in\n∗osunaruiz.david@usal.esthe domains between DWs can be efficiently driven by\npure electrical means. Both UD and DU DWs move\nwith the same velocity along straight stacks, but some\nimplementations of these memory or logic devices would\nrequire to design 2D circuits, where straight parts of\nHM/FM/Ox stack are connected each other with curved\nor semi-rings sections. However, recent experimental ob-\nservations [10] and theoretical studies [11] have pointed\nout that adjacent UD and DU DWs move with differ-\nent velocity along curved HM/FM/Ox stacks, which is\ndetrimental for applications because the size of the do-\nmain between adjacent DWs changes during the motion,\nwith the perturbation of the information coded therein.\nTherefore, other systems must be proposed in order to\ndesign reliable 2D circuits for DW-based memory and\nlogic devices.\nOther stacks with materials and/or layers with anti-\nferromagnetic coupling, such as synthetic antiferromag-\nnets (SAF) and ferrimagnetic (FiM), have proven to out-\nperform FM in terms of current-driven DW dynamics\n[11–15]. Ultrafast magnetization dynamics in the THz\nregime, marginal stray field effects and insensitivity to\nexternal magnetic fields are other significant advantages\nof materials with antiferromagnetic coupling with respect\nto their FM counterparts. As conventional antiferromag-\nnets (AFs), FiM alloys are also constituted by two spec-\nimens, typically a rare earth (RE) and transition metal\n(TM), that form two ferromagnetic sublattices antiferro-\nmagnetically coupled to each other. GdFeCo, GdFe or\nTbCo are archetypal FiM alloys, with the RE being Gd\nor Tb and the TM being FeCo or Co. In contrast to AFs\nwith zero net magnetization, the magnetic properties of\nFiMs, such as magnetization and coercivity, are largely\ninfluenced by the relative RE and TM composition (or\nequivalently, temperature). This fact offers additional\ndegrees of freedom to control the current-driven DW ve-\nlocity. The spontaneous magnetization of each sublattice\nMS,ican be tuned by changing the composition of the\nFiM and/or the temperature of the ambient ( T) [13, 14].\nFor a given composition of the FiM (RE xTM 1−x), therearXiv:2401.13617v1 [cond-mat.mtrl-sci] 24 Jan 20242\nare two relevant temperatures below the Curie thresh-\nold. One is the magnetization compensation temperature\n(TM) at which the saturation magnetization of the two\nsublattices are equal ( Ms1(TM) =Ms2(TM)), so the FiM\nbehaves as a perfect antiferromagnetic material, with\nzero net magnetization and diverging coercive field. The\nother is the temperature at which the angular momentum\ncompensates, TA, at which Ms1(TA)/γ1=Ms2(TA)/γ2,\nwhere γiis the gyromagnetic ratio of each sublattice\n(i:1,2 for 1:TM and 2:RE). As the gyromagnetic ratio\ndepends on the Land´ e factors ( gi) which are different for\neach sublattice, the angular compensation temperature\nTAis in general different from the magnetization compen-\nsation temperature ( TM). Consequently, the FiM have\na net magnetization at TA, so conventional techniques\nused for FMs can be also adopted to detect the magnetic\nstate of FiM samples [16]. Moreover, recent experimen-\ntal observations have evidenced that the current-driven\nDW velocity along straight HM/FiM stacks can be sig-\nnificantly optimized at the angular momentum compen-\nsation temperature ( T=TA), with velocities reaching\nVDW∼2000 m/s for typical injected density current of\nJHM∼1 TA/m2along the HM underneath [14]. The\nDW velocity drops either below ( T < T A) and above\n(T > T A) angular momentum compensation. Note that\nalternatively to tuning the temperature for a fixed com-\nposition xof the FiM alloy RE xTM 1−x, even working at\nroom temperature ( T= 300 K) the DW velocity peaks\nat a given composition where angular momentum com-\npensates [13]. Therefore, both studies, either fixing the\ncomposition ( x) and changing temperature of the ambi-\nent (T), or fixing the ambient temperature and modifying\nthe FiM composition are equivalent for our purposes of\nDW dynamic. Although the current-driven DW motion\n(CDDWM) along HM/FiM stacks suggests their poten-\ntial for memory and logic applications, previous studies\nhave been mainly focused on straight FiM strips [13–15].\nThe further develop of novel DW-based devices also re-\nquires to analyze the dynamics of DWs along HM/FiM\nwith curved parts which would connect straight paths to\ndesign any 2D circuit. Such investigation of the dynam-\nics along curved is still missing, and it is the aim of the\npresent study.\nHere we theoretically explore the CDDWM along\ncurved HM/FiM stacks by means of micromagnetic ( µm)\nsimulations. Our modeling allows us to account for the\nmagnetization dynamics in the two sublattices indepen-\ndently. We explore the CDDWM below, at and above\nthe angular momentum compensation (AMC) for differ-\nent curved samples, with different widths and curvatures,\nand considering the realistic spatial distribution of the\ninjected current along the HM. In particular, we will in-\nfer and isolate the relevance of different aspects govern-\ning such dynamics, as the role of the non-uniform cur-\nrent and other purely geometrical aspects of the curved\nshape. This work completes previous studies on straight\nsamples [11, 13–15], and will be practical for designing\nmore compact and efficient DW-based devices. The restof the paper is organized as follows. In Sec. II we de-\nscribe the numerical details of the micromagnetic model\nalong with the material parameters and the geometri-\ncal details of the evaluated samples. Sec. III presents\nthe micromagnetic results of the CDDWM in different\nscenarios. Firstly, exploring the role of the FiM sam-\nple width ( w) for a fixed the curvature ( ρ, given by the\ninverse of the average radius, ρ= 1/re), and secondly\nfixing the width and varying the curvature. After that,\nwe present results which allow us to infer the role of non-\nuniform current and geometrical aspect ( w, ρ) comparing\ncurved and straight samples. The main conclusions are\nsummarized in Sec. IV.\nII. MATERIALS AND METHODS\nCDDWM is numerically studied here along curved\nHM/FiM stacks as schematically shown in Fig. 1, where\nriandroare the inner and outer radius rorespectively,\nandre= (ro+ri)/2 is the mean effective radius. wand\ntFiMare the width and the thickness of the FiM respec-\ntively. The relaxed magnetization configuration of the\nsublattice i= 1, shown in Fig. 1 (opposite configuration\nin sublattice i= 2), serves as the initial state to study\nthe CDDWM upon of current injection along the HM un-\nderneath. The temporal evolution of the magnetization\nof each sublattice is given by the Landau-Lifshitz-Gilbert\nequation (LLG) [17],\nd⃗ mi(t)\ndt=−γ0,i⃗ mi(t)×⃗Heff,i+αi⃗ mi×d⃗ mi(t)\ndt+⃗ τSOT,i ,\n(1)\nwhere here the sub-index istands for i: 1 and 2 sub-\nlattices respectively. γi=giµB/ℏandαiare the gy-\nromagnetic ratios and the Gilbert damping constants,\nrespectively. giis the Land´ e factor of each layer, and\n⃗ mi(⃗ r, t) =⃗Mi/Ms,iis the normalized local magnetization\nto its saturation value ( Ms,i), defined differently for each\nsublattice: Ms,i(i: 1,2). In our micromagnetic model the\nFiM strip is formed by computational elementary cells,\nand within each cell we have two magnetic moments, one\nfor each component of the FiM. The respective effective\nfield ( ⃗Heff,i) acts on the local magnetization of each sub-\nlattice ( ⃗ mi(⃗ r, t)), and it is the sum of the magnetostatic,\nthe anisotropy (PMA), the DMI and the exchange fields\n[11, 15]. The magnetostatic field on each local moment\nin the sublattice is numerically computed from the av-\nerage magnetization of each elementary cell using simi-\nlar numerical techniques as for the single FM case (see\n[11, 15]). We checked that the demagnetising field has a\nmarginal influence in the simulation results compared to\nother contributions to the effective field. For the PMA\nfield, the easy axis is along the out-of-plane direction ( z-\naxis), and the anisotropy constants for each sublattice are\nKu,i(PMA constant). Diis the DMI parameter for each\nsublattice i: 1,2 [11, 15]. The exchange field of each sub-\nlattice includes the interaction with itself (intra-lattice3\nexchange interaction, ⃗Hexch,i ) and with the other sublat-\ntice (inter-lattice exchange interaction, ⃗Hexch, 12). The\ninter-lattice exchange effective field is computed as for\na single FM sample, Hexch,i =2Ai\nµ0Ms,i∇2⃗ mi, where Aiis\nthe intralattice exchange parameter. The inter-lattice ex-\nchange contribution ⃗Hexch, 12to the effective field ⃗Heff,i,\nacting on each sublattice is computed from the corre-\nsponding energy density, ωexch,i =−Bij⃗ mi·⃗ mj, where\nBij(in [J m−3]) is a parameter describing the inter-lattice\nexchange coupling between sublattices (here, we used the\nnotation i: 1 and j: 2).\nIn Eq.(1), ⃗ τSOT,i are the SOTs acting on each sub-\nlattice, which are related to the electrical current along\nthe HM ( ⃗JHM). Based on preliminary studies [18], here\nwe assume that ⃗ τSOT,i is dominated by the spin Hall ef-\nfect (SHE), so ⃗ τSOT,i =−γ0HSL⃗ mi×(⃗ mi×⃗ σ) where\nHSL=ℏθSH,iJHM\n2|e|µ0MstFiM[19]. ℏis the Planck constant, and\nθSH,i is the spin Hall angle, which determines the ra-\ntio between the electric current and the spin current\n(Js=θSHJHM) for each sublattice. ⃗ σ=⃗ uJ×⃗ uzis\nthe unit vector along the polarization direction of the\nspin current generated by the SHE in the HM, being\northogonal to both the direction of the electric current\n⃗ uJand the vector ⃗ uzstanding for the normal to the\nHM/FiM interface. For a longitudinal current ( ⃗ uJ=⃗ ux),\nthe spin current is polarized along the transverse di-\nrection, ⃗ σ=−⃗ uy. For curved samples where the cur-\nrent density ⃗JHM=JHM(r)⃗ uJhas azimuthal direction\n(⃗ uJ=−⃗ uϕ), the direction of the polarization is radial,\n⃗ σ=⃗ uJ×⃗ uz=⃗ ur, as shown in Fig 1. A potential dif-\nference is applied between the ends of the curved track\nto inject current in the right circulation. Therefore, a\ngap of 25 nm is also modelled, leading to a split ring\nshape for the strip (see inset in Fig. 1). The spatial\ndistribution of current as a function of the radial coor-\ndinate ( ri< r < r o) is taken from [11, 20], and it de-\npends on the width ( w) and the radial distance ( r) as\nJHM(r) =J0w/(rlog (1 + w/ri)), where J0is the nom-\ninal, uniform current density, in an equivalent straight\nstrip of same cross-section ( w×tHM, where tHMis the\nthickness of the HM strip).\nIn order to illustrate the current-driven DW dynam-\nics along curved HM/FiM stacks we fix tFiM = 6 nm,\nand samples with different widths ( w) and radii ( re)\nwere evaluated. The following common material pa-\nrameters were adopted for the two sublattices i: 1,2:\nAi= 70 pJ/m, Ku,i= 1.4×106J/m3,αi= 0.02,\nDi= 0.12 J/m2,θSH,i = 0.155. The strength of the\nantiferromagnetic coupling between the sublattices was\nfixed to Bij≡B12=−0.9×107J/m3. The gyro-\nmagnetic ratios ( γi=giµB/ℏ) are different due to the\ndifferent Land´ e factor: g1= 2.05 and g2= 2.0. The\nsaturation magnetization of each sublattice Ms,ican be\ntuned with the composition of the FiM and/or with the\ntemperature of the ambient ( T). Here, we assume the\nfollowing temperature dependences for each sublattice:Ms,i(T) =Ms,i(0)\u0010\n1−T\nTC\u0011ai\n, where TC= 450 K is the\nCurie temperature of the FiM, Ms,1(0) = 1 .4×106A/m\nandMs,2(0) = 1 .71×106A/m are the saturation magne-\ntization at zero temperature, and a1= 0.5 and a2= 0.76\nare the exponents describing the temperature depen-\ndence of the saturation magnetization of each sublattice.\nThe temperature at which the net saturation magneti-\nzation vanishes ( Ms,1(TM) =Ms,2(TM)) is TM= 241 .5\nK, and the angular momentum compensation temper-\nature corresponding to Ms,1(TA)/g1=Ms,2(TA)/g2, is\nTA= 260 K. We evaluate the CDDWM below, at and\nabove the angular momentum compensation adopting\nthree representative temperatures: T= 220 K < T A,\nT= 260 K = TAandT= 300 K > TA. Samples were\ndiscretized using a 2D finite difference scheme using com-\nputational cells with ∆ x= ∆y= 0.2 nm and ∆ z=tFiM.\nSeveral tests were carried to certify that the presented re-\nsults are free of discretization errors.\nFIG. 1. Scheme showing the relaxed states of spins in sub-\nlattice i= 1, in the positive z-direction (white domain), in\nthe negative z-direction (black domain) and in the plane of\nthe strip for an ‘Up to Down’ (UD) domain wall (purple)\naccording to the current direction, for an exemplary curved\nstrip. The direction of the applied electric current (red arrow)\nin the Heavy Metal beneath the magnetic strip, generated\nfrom a potential difference ∆ V(see inset), is shown as well\nas the geometrical parameters of the strip.\nIII. MICROMAGNETIC RESULTS\nDue to the several combination of parameters to con-\nsider in our study, we divided this section in three sub-\nsections: (A) The study on the influence of the strip\nwidth ( w); (B) the study on curvature ( ρ=r−1\ne); and (C)\nsame studies for a straight strip with identical material\nparameters, to explore by comparison the effects of curva-\nture on the DW dynamics. In parts (A) and (B), scenar-\nios for three different temperatures, T1= 220 K , T2= 2604\nK, T3= 300 K are considered, to study the DW motion\nbelow the AMC ( T1), at the AMC ( T2) and above the\nAMC ( T3). We also define and refer to T3= 300 K as for\n‘room temperature’ in our study. Note that a change in\ntemperature only affects MSin our model, therefore it\nhas equivalent effects to changing material composition\n[13]. In addition to DW velocity, we also characterize the\ninertial motion of the DW as a function of current den-\nsity. As an example, Fig. 2 shows typical results of the\nDW position and its velocity in a ring-like strip ( w= 256\nnm and re= 384 nm) under a density current JHM= 2\nTA/m2and at T= 260 K. Qualitatively similar results\nare obtained at T= 220 K and T= 300 K (not shown).\nInsets show the (clockwise) DW displacement as a func-\ntion of time for one sublattice ( i= 1).\nA. Influence of width for a fixed curvature\nIn this study, the curvature is fixed ( re= 384 nm) and\nwidth ( w) is varied from 56 nm to 296 nm in steps of 40\nnm. Fig. 3 shows the results for the terminal DW veloc-\nity (|VDW,i|) as a function of the nominal density current\nJHM=j0, equivalent to the homogeneous density cur-\nrent in a straight strip with the same cross-section. In the\nnext sections, we use the notation JHM= J for simplic-\nity. Fig. 3 shows that temperature has a noticeable effect\non the terminal velocity on the DW type equally, Up to\nDown domain (UD) or Down to Up domain (DU). As it\nwas expected from previous work on straight FiM strips\n[11, 13], at TAthe DW velocities are greater. Also, the\nDW velocities increase for narrower strips (red symbols).\nIn fact, the observed trend is very similar to straight\nstrips: the terminal velocity is maximum at the temper-\nature of AMC, TA= 260 K and significantly increased,\nexceeding 2000 m/s for the narrowest strip as compared\nto∼1800 m/s for the widest. These results also suggest\nthat the DWs velocities are equal for both types of DWs\n(UD and DU), which would lead to no distortion of the\nsize of a domain between two adjacent DWs travelling\nalong the curved strip. This result is significantly differ-\nent from FM systems [11].\nAtT̸=TA, the DW velocity is reduced either increas-\ning or reducing temperature with respect to TA, leading\nto velocities around 1100 m/s, generally regardless of the\nwidth and DW type. For a given Jvalue, as the strip\ngets wider, however, the velocity is slightly smaller but\nthese differences are negligible (see Fig. 3(a), (c), (d),\n(f)). This result contrasts with that of a FM strip, where\na greater difference of velocities between a DU and a UD\nalong a curved strip was shown [11].\nTo characterize the inertial motion of the DW, we eval-\nuate the temporal evolution of the DW velocity ( V(t),\ncomputed from the spatial averaging of m1,z(t)) as a\nfunction of time (or instant velocity) under a current\nsquare pulse of duration 0.1 ns and start at t= 0. The\nµm results can be fitted to the following exponentials:\nV∞(1−e−t/τr) during the duration of the pulse ( t≤0.1\nxf\n0.00 ns\n0.50 ns\n0.13 ns\n0.25 ns\n0.38 ns\n(a)\n(b)\n(c)(d)FIG. 2. Micromagnetic results of applying a uniform JHM=\nJ0= 2 TA/m2(a) showing the relative and final position xf\n(b) and velocity VDW(c), as a function of time t, of an UD\nDW in a curved strip ( w= 256 nm and re= 384 nm), at\nT= 260 K. Insets (d) are snapshots of the magnetic config-\nuration in sublattice i= 1 at different times (highlighted by\nthe vertical dotted lines in (a-c)). Green solid lines are for\nguiding the eye and are co-parallel with the strip radius.\nns), and V∞e−t/τf, after the pulse ends ( t > 0.1 ns),\nwhere V∞is the DW terminal velocity (see Fig. 2(a)-(f)).\nThe characteristic relaxation times τr(or rising time) and\nτf(or fall time) represent the duration of such transients\nand characterize the inertial motion of the DW. These\nparameters can be extracted by fitting the µm results to\nthe exponentials (see solid curves in Fig. 4(a)).\nAlthough simulations were performed for both types of\nDWs, note that we only present here results for the DU\ntype wall, for sake of simplicity. Identical results (not\nshown) were obtained for the UD DW. Fig. 4(a) show\nthe ‘instantaneous’ DW velocity V(t) and the relaxation\ntimes τfor three selected values of J (see solid symbols) at\nT=TAfor the widest strip ( w= 296 nm). Solid lines are5\n𝑤\n260 K 260 K220 K 220 K\n300 K 300 K(a)\n(b)\n(c)(d)\n(e)\n(f)\nT=220 K\nT=260 K\nT=300 K\nDUUD\n(g)\nw=296 nm\nww (nm)\n56 \n176 \n296 \nw=56 nmw (nm)\n56 \n176 \n296 UD\nDU\nT=220 K\nT=260 K\nT=300 K\nFIG. 3. Terminal velocities as a function of J for a UD (a-c)\nand a DU (d-f) DW obtained for sub-lattice i= 1 and for\nthree different temperatures: below, above and at the AMC\ntemperature (220 K, 300 K and 260 K, respectively). Strips\nfor the two limiting cases are shown in the red and blue con-\ntour insets at the top. (g) Terminal velocities as a function\nofwfor a UD (full symbols) and a DU (open symbols) type\nwall for J = 2 .35 TA/m2and the three chosen temperatures.\nInset in (g) shows J(r) for two values of w. Red dashed line\nindicates J = 2 .35 TA/m2.\nthe exponential curves to which the obtained simulated\ndata is fit. For each current, the minimal τis expected\nforT=TA= 260 K. Fig. 4(b) shows that τrandτffor\nthe two limiting cases ( w= 56 nm and w= 296 nm) are\nquantitatively similar, in the order of 0.02 ns, since they\nfall within the 95% confidence interval, set by the largest\nerror bars obtained for τfrom the fitted results, among\nall J. Also, all values are similar in order to the step-size\nJ = 2.35 TA/m2\nJ = 1.60 TA/m2\nJ = 0.85 TA/m2T = 260 K\n(a) (b)FIG. 4. Instantaneous DW velocities V(t) for three values of\nJ and for a curved strip of re= 384 nm and w= 296 nm at\nT=TA(a). Symbols are the µm data, from which τ(rise and\nfall times) are extracted for the narrowest and widest strips\n(b). Dashed lines in (b) are the upper and lower bounds of a\n95% confidence interval.\nused in simulations, 0.01 ns (see Fig. 4(d)).\nSimilar values of τwere obtained for strips of other\nwidths. Relaxation times are not noticeably influenced\nby temperature, and they generally remain within the\nrange of 0.01 ∼0.03 ns for T= 200 K and T= 300 K.\nThis is more than one order of magnitude smaller than\nin FM strips, the latter being about ∼1 ns according to\nRef.[21]. Besides, the relaxation times of current-driven\nDWs in curved strips found here are in good agreement\nwith those from field-driven or thermally driven DWs in\nantiferromagnetic straight strips, in the order of picosec-\nonds [22, 23].\nB. Influence of curvature for a fixed width\nIn this study, the strip width is fixed to w= 256 nm\nand the curvature parameter ρis varied. In other words,\nthe equivalent radii re(re=ρ−1) is varied from 134\nnm to 534 nm in steps of 50 nm. Fig. 5 shows the re-\nsults for the terminal velocity ( |VDW, 1|) of DU and UD\nDWs, for several values of rein nanometers, where the\nred (blue) curve corresponds to the smaller (greater) val-\nues, for three different temperatures.\nFig. 5(a)-(f) shows that, at T=TAand for a given\ncurvature, the DW velocities of DU and UD types are\nvery similar for the whole range of currents explored.\nDW velocity reduces as the curvature increases (see Fig.\n5(g)). It is worth noting that the latter cannot be a con-\nsequence of only a nonuniform J(r) as defined in [11]. In\nfact, for curved-most strips ( re= 134 nm), the spatial-\ndependent density current J(r) varies with rsimilarly as\nit does for changing w(see inset in Fig. 5(g) and in Fig.\n3(g)), which would suggest similar variations to DW ve-\nlocities as those found in Fig. 3(g). In other words, the\nimpact of the non-uniform J(r) is not so relevant to be\nthe only source of the big differences between the DW\nvelocities for large and small curvatures (orange symbols\nin Fig. 5(g) for re= 134 nm and re= 484 nm, respec-\ntively). Fig. 5(g) also shows that as the strip curvature6\n260 K 260 K220 K 220 K\n300 K 300 K(a)\n(b)\n(c)(d)\n(e)\n(f)\n𝑟𝑒\n(g)re (nm)\n134 \n384 \n584 \nre (nm)\n134 \n384 \n584 UD\nDU\nT=220 K\nT=260 K\nT=300 K\nDUUD\nT=220 K\nT=260 K\nT=300 Kre = 134 nm\nre = 534 nm\nFIG. 5. Terminal velocities as a function of J for a UD (a-c)\nand a DU (d-f) DW in the curved strip obtained for sub-\nlattice i= 1 and for three different temperatures: below,\nabove and at the AMC temperature (220 K, 300 K and 260\nK, respectively). Strips for the two limiting cases are shown\nin the red and blue contour insets at the top. (g) Terminal\nvelocities as a function of refor a UD (full symbols) and a DU\n(open symbols) type wall for J = 2 .35 TA/m2and the three\nchosen temperatures. Inset in (g) shows J(r) for two values\nofre. Red dashed line indicates J = 2 .35 TA/m2.\nis reduced, DW velocity converges to the straight strip\ncase ( re→ ∞ ).\nForT̸=TA, the dependence of the DW velocity with\ntemperature is minimal regardless of the DW type. As\nthe strip curvature increases there is a prominent change\nin the maximal terminal DW velocity for both DW types.\nHowever, the relative difference of velocities is almost\nJ = 2.35 TA/m2\nJ = 1.60 TA/m2\nJ = 0.85 TA/m2\n(a) (b)FIG. 6. Instantaneous DW velocities V(t) for three values of\nJ and for a curved strip of re= 584 nm and w= 256 nm at\nT=TA(a). Symbols are the µm data, from which τ(rise and\nfall times) are extracted for the narrowest and widest strips\n(b). Dashed lines in (b) are the upper and lower bounds of a\n95% confidence interval.\nnegligible. Therefore, results suggest that the strip cur-\nvature affects in a similar way to width, and equally, to\nboth DWs. In other words, the terminal velocity is signif-\nicantly reduced as curvature (or width) increases, while\nthe differences between DWs remain negligible (see Fig.\n5(g)). This behavior is even more pronounced at T=TA.\nAs discussed in section III.A, the latter would imply that\nthe robustness of a transmitted bit, encoded in a domain\nbetween two DWs, can be optimised in such curved-most\nstrips and reaches larger velocities in the strip.\nFig. 6(a) show the DW velocity as a function of time\nand for three selected values of J for an effective radius of\nre= 534 nm (least curved strip) and intermediate width\nw= 256 nm, at T=TA. Results look quantitatively sim-\nilar to those shown in Fig. 4(a), where rewas fixed to an\nintermediate value of 384 nm. As in Fig. 4(b), Fig. 6(b)\nshows that τrandτffor the two limiting cases ( re= 134\nnm and re= 534 nm) are quantitatively similar, in the\norder of 0.02 ns. For all the FiM curved strips explored\nat, above and below AMC, τrandτfremain within the\nrange of 0.01 ∼0.03 ns, approximately one order of mag-\nnitude less than their FM counterparts. This is in good\nagreement with results presented in the previous section\nand other work in straight strips [8], which further sup-\nports the negligible inertia of DWs in such FiM systems.\nC. Discussion on the effective influence of a curved\nshape on the wall velocity\nA non-uniform current distribution is expected to in-\nfluence the terminal velocity of the DW for a given cur-\nvature, specially for wide curved strips [11]. In this sec-\ntion, to explore further the degree of influence of the\nnon-uniform current, equivalent studies on wandρon\na straight strip with an artificially implemented non-\nuniform J( r=y) atT=TAare performed. A straight\nstrip is a bounding case for a curved strip that shows no\neffective curvature ( re→ ∞ , ρ→0) and an homogeneous\ndensity current J=J0. Therefore, we explore whether7\n(c)(a)\n(d)w = 256 nm re = 384 nm(b)\nre = 384 nm w = 256 nm\nJ = Jx(y, re)\nJx = J 0J = Jx(y, w)\nJx = J 0\nJ = Jφ(r, re)J = Jφ(r, w)\nJφ = J 0\nFIG. 7. DW terminal velocities in a straight strip for an in-\nhomogeneous J(y, w, r e) (blue circles) and for a uniform J0\n(black crosses) as a function of radius re(a) and width w(b)\nat same temperature ( T= 260 K). Insets in (a) show the mag-\nnetic configuration of sublattice i= 1 at t= 0 and schematics\nof the current spatial distribution in the strips as examples.\n(b) DW terminal velocities in a curved strips with the same\nparameters as a function of radius re(c) and width w(d). As\nan example, inset in (c) shows the radial dependence of an in-\nhomogenous current distribution in such a strip. Dotted lines\nhighlight the cases where the two geometrical parameters ( w\nandre) are coincident among all the studies.\n‘curvature ( ρ) effects’ are mainly dominated by the in-\ntrinsic inhomogenous current, or whether they can also\narise from the curved shape itself [24]. We aim to discern\nthe actual influence of an inhomogeneous current, as part\nof an more global effect due to the curved shape. For the\nfollowing study, and since the straight shape must be re-\ntained, re(or equivalently ρ) is artificially modified in the\nnon-uniform current distribution expression: ⃗J(y, w, r e)\n[11] in the x-direction, as if the strip was curved. Note\nthat ρrepresents the inverse of the averaged or effective\ncurvature radius of the strip ( ρ=r−1\ne) and not the cylin-\ndrical radial coordinate ( r) in the system.\nFig. 7(a-b) show the velocity of an UD wall in the\nstraight strip for J=J0= 2.35 TA/m2(black crosses)\nand for an inhomogeneous J(y) (blue circles), varying re\ninJ(y, re) (see insets) for a fixed width w= 256 nm, and\nvarying width ( w) for a fixed re,J(y, re= 384). While it\nis expected that an inhomogeneous J(y, w, r e) will influ-\nence the DW velocity (especially for curved-most strips,\nseere= 134 nm in (a)), the differences with the case of\nJ=J0are almost negligible. Fig. 7(c-d) show resultsfor curved strips. For these cases, wandreare naturally\nvaried in J(r, w, r e) by modifying the shape itself. From\nthe standpoint of the applied current, this is expected\nto be equivalent to doing it by changing the shape itself.\nResults from an inhomogeneous current (blue circles, re-\nproduced from Fig. 3(g) and 5(g)) consistently tend to\nconverge to the straight strip as ρis reduced. The strip\nis straight when ρ= 0 ( re→ ∞ ).\nWhen the strip is either straight (a-b) or curved (c-\nd), for both studies (fixing wand varying reor vice-\nversa), the DW velocity is found to be the same when\nthe geometrical parameters are coincident, i.e., w= 256\nnm and re= 384 nm, as expected (see horizontal dot-\nted lines in (a-b) or (c-d)). However, when the shape\nis different, even in the cases when wandreare coinci-\ndent (and therefore, J(w, re) is expected to also be the\nsame), different DW velocities are obtained (see vertical\ndotted lines in (a-c) and (b-d)) below and slightly above\n2000 m/s, respectively. Moreover, by modifying either re\norwin the curved strip by directly changing its shape\n(see (c-d) and previous sections), there is a clear larger\nimpact on DW velocity, than by artificially (but equiva-\nlently) modifying reorwin the straight strip (see (a-c)).\nIn the curved strips, the trend when varying reorwis\nqualitatively similar regardless to the homogeneity of the\napplied J(see black crosses and blue circles in either (c)\nor (d)). Since J( r) is modeled in an equivalent way in all\nstudies by simply changing J( r) = J( y) for the straight\nstrip, marked differences between the wall velocity in (a-\nb) (straight strip) and (c-d) (curved strip), specially for\nvery curved strips, suggest that not only the non-uniform\nJ(r) is influencing the UD DW motion.\nOur results suggest that the curved shape may have\nan intrinsic influence on the wall velocity, manifested as\na more marked dependence with wandre, regardless of\nthe inhomogeneity of the current density (see Fig.(c-d)),\nas the shape becomes more curved. An influence due to\nthe inhomogeneous current, still appears naturally in the\ncurved strip, but may have a lesser impact compared to\nother geometrical factors (see differences between black\ncrosses and blue circles).\nIV. DISCUSSION AND SUMMARY\nWe have provided a study on DW motion in curved\nFiM strips, particularised to one of the two strongly cou-\npled sub-lattices, for three different temperatures, and as\na function of geometrical parameters for a HM/FiM/Ox\nmultilayer structure. We observe an absence of tilting of\nthe DW and domain distortion at different temperatures,\n40 K above and below the angular momentum compen-\nsation temperature.\nWidth and curvature effects on the DW velocity are\ndiscussed. Besides contributions from a non-uniform\nJ(r), there is an overall significant influence from the\nshape of the strip itself on DW velocity. This implies\nthat, for a fixed temperature (or composition), DW ve-8\nlocity can be optimised by optimising the geometrical\nparameters of the curved strip. The relative differences\nbetween a DU and a UD walls are marginal in general.\nIn other words, geometrical factors affect them almost\nequally, which is positive for a robust transmission of a\nbit encoded in an Up or Down domain between two adja-\ncent DWs. With reducing current, differences in veloci-\nties between curved and straight strips are still minimised\nat the expense of slower DWs. Similar effect is observed\nas width or curvature is increased. This is beneficial for\ndesigning intricate 2D circuit tracks combining curved\nand straight sections, while preserving DW velocities still\nlarger than those found in their FM counterparts. Also,\nDWs in a curved FiM strip show a negligible inertia in\ncontrast to their FM counterparts ( τFiM << τ FM) for\nall the explored scenarios. The DWs start to move and\nstop almost immediately ( τFiM∼0.02 ns) after the ap-\nplication or removal of current.\nConsidering the obtained results altogether and assum-\ningT̸=TA, which will be most of the experimental cases\nat room temperature ( T= 300 K), our study allows us\nto conclude that narrow enough FiM strips are ideal can-\ndidates for designing curved tracks for 2D spintronic cir-\ncuits of an arbitrary shape based on CDDWM, where bits\nare encoded in domains separated by walls. This is dueto very fast rise and fall times ( τr∼τf<<0.1 ns), high\nvelocities ( VDW>1000 m/s) and negligible distortion of\nthe two types of DWs (UD and DU) in all the scenarios\nexplored in this work. Greater DW terminal velocities\nand smaller time responses in curved FiM strips than\nthose in their FM counterparts are obtained. These re-\nsults can help in the further research, development and\nimprovement of FiM-based spintronic circuitry that may\nrequire compactness and high-speed functionality with\nhigh robustness to DW (and/or domain) distortion.\nV. ACKNOWLEDGEMENTS\nThis work was supported by project SA114P20 from\nJunta de Castilla y Leon (JCyL), and partially supported\nby projects SA299P18 from JCyL, MAT2017-87072-C4-\n1-P and PID2020-117024GB-C41 from the Ministry of\nEconomy, Spanish government, and MAGNEFI, from\nthe European Commission (European Union). All data\ncreated during this research are openly available from\nthe University of Salamanca’s institutional repository at\nhttps://gredos.usal.es/handle/10366/138189\n[1] A. Hubert and R. Sch¨ afer, in Magnetic Domains: The\nAnalysis of Magnetic Microstructures (Springer-Verlag\nBerlin Heidelberg, 1998).\n[2] S. S. P. Parkin, “Shiftable magnetic shift register and\nmethod of using the same,” (US6834005B1, 2004).\n[3] S. Parkin and S.-H. Yang, Nature nanotechnology 10,\n195—198 (2015).\n[4] I. Miron, T. Moore, H. Szambolics, L. Buda-Prejbeanu,\nS. Auffret, B. Rodmacq, S. Pizzini, J. Vogel, M. Bonfim,\nA. Schuhl, and G. Gaudin, Nature materials 10, 419\n(2011).\n[5] S. Emori, U. Bauer, S.-M. Ahn, E. Martinez, and\nG. S. D. Beach, Nature materials 12, 611—616 (2013).\n[6] P. Haazen, E. Mur´ e, J. Franken, R. Lavrijsen,\nH. Swagten, and B. Koopmans, Nature Materials 12,\n299 (2013).\n[7] K.-S. Ryu, L. Thomas, S.-H. Yang, and S. Parkin, Na-\nture nanotechnology 8(2013), 10.1038/nnano.2013.102.\n[8] J. Torrejon, E. Martinez, and M. Hayashi, Nature com-\nmunications 7, 13533 (2016).\n[9] I. Dzyaloshinsky, Journal of Physics and Chemistry of\nSolids 4, 241 (1958).\n[10] C. Garg, S.-H. Yang, T. Phung, A. Pushp, and S. Parkin,\nScience Advances 3, e1602804 (2017).\n[11] O. Alejos, V. Raposo, and E. Mart´ ınez, “Domain wall\nmotion in magnetic nanostrips,” in Materials Science and\nTechnology (American Cancer Society, 2020) pp. 1–49.\n[12] S.-H. Yang, K.-S. Ryu, and S. Parkin, Nature nanotech-\nnology 10(2015), 10.1038/nnano.2014.324.[13] S. A. Siddiqui, J. Han, J. T. Finley, C. A. Ross, and\nL. Liu, Phys. Rev. Lett. 121, 057701 (2018).\n[14] L. Caretta, M. Mann, F. B¨ uttner, K. Ueda, B. Pfau,\nC. G¨ unther, P. Hessing, A. Churikova, C. Klose,\nM. Schneider, D. Engel, C. Marcus, D. Bono,\nK. Bagschik, S. Eisebitt, and G. Beach, Nature Nan-\notechnology 13, 1154 (2018).\n[15] E. Mart´ ınez, V. Raposo, and ´Oscar Alejos, Journal of\nMagnetism and Magnetic Materials 491, 165545 (2019).\n[16] S. Arpaci, V. Lopez-Dominguez, J. Shi, L. S´ anchez-\nTejerina, F. Garesci, C. Wang, X. Yan, V. K. Sang-\nwan, M. A. Grayson, M. C. Hersam, G. Finocchio and\nP. Khalili Amiri, Nature Communications 12(2021),\n10.1038/s41467-021-24237-y.\n[17] L. Landau and E. Lifshitz, Phys. Zeitsch. der Sow. 8,\n153–169 (1935).\n[18] E. Martinez, N. Perez, L. Torres, S. Emori, and\nG. S. D. Beach, Journal of Applied Physics 115(2014),\n10.1063/1.4881778.\n[19] J. C. Slonczewski, Journal of Magnetism and Magnetic\nMaterials 159, L1 (1996).\n[20] O. Alejos, V. Raposo, L. Sanchez-Tejerina, R. Tomasello,\nG. Finocchio, and E. Martinez, Journal of Applied\nPhysics 123(2018).\n[21] A. Thiaville, Y. Nakatani, F. Pi´ echon, J. Miltat, and\nT. Ono, European Physical Journal B 60, 15 (2007).\n[22] O. Gomonay, T. Jungwirth, and J. Sinova, Phys. Rev.\nLett. 117, 017202 (2016).\n[23] S. Selzer, U. Atxitia, U. Ritzmann, D. Hinzke, and\nU. Nowak, Phys. Rev. Lett. 117, 107201 (2016).\n[24] L. Yang, East Asian Journal on Applied Mathematics 7,\n837–851 (2017)." }, { "title": "2201.09067v1.Universal_criteria_for_single_femtosecond_pulse_ultrafast_magnetization_switching_in_ferrimagnets.pdf", "content": "Universal criteria for single femtosecond pulse ultrafast magnetization switching in\nferrimagnets\nF. Jakobs and U. Atxitia\u0003\nDahlem Center for Complex Quantum Systems and Fachbereich Physik,\nFreie Universität Berlin, 14195 Berlin, Germany\nSingle-pulse switching has been experimentally demonstrated in ferrimagnetic GdFeCo and\nMn2RuxGa alloys. Complete understanding of single-pulse switching is missing due to the lack\nof an established theory accurately describing the transition to the non-equilibrium reversal path\ninduced by femtosecond laser photo-excitation. In this work we present general macroscopic theory\nfor the magnetization dynamics of ferrimagnetic materials upon femtosecond laser excitation. Our\ntheory reproduces quantitatively all stages of the switching process observed in experiments. We\ndirectlycompareourtheorytocomputersimulationsusingatomisticspindynamicsmethodsforboth\nGdFeCo and Mn 2RuxGa alloys. We provide explicit expressions for the magnetization relaxation\nrates in terms of microscopic parameters which allows us to propose universal criteria for switching\nin ferrimagnets.\nUltrafast magnetization switching induced by a single\nfemtosecond laser pulse has attracted a lot of attention\nas a promising solution for low energy, faster memory\napplications. [1–15]. Until recently only GdFeCo\nalloys and synthetic ferrimagnets [8], presented the\nability to switch under either optical femtosecond laser-\n[3, 4, 8] or electric picosecond current-pulses [16,\n17]. Although several micro- and macroscopic models\nhave reproduced single-pulse switching in GdFeCo\nferrimagnets [18–33], complete understanding of the\nrole of electrons, lattice and spin sublattices and their\nmutual interactions remains a challenge [34]. The\nexisting criteria for switching rely on the existence of\ntwo antiferromagnetically coupled magnetic sublattices\nshowingdistinctdynamicalresponsetofemtosecondlaser\nphoto-excitation. While in single species ferromagnets\nsuch as 3d transition metals, relaxation of angular\nmomentum occurs via dissipation into other degrees\nof freedom – relativistic relaxation – in two-sublattice\nmagnets, additionally, relaxation can occur via angular\nmomentum exchange between sublattices – exchange\nrelaxation. By driving the spin system into a non-\nequilibrium state where exchange relaxation dominates,\na non-equilibrium ultrafast reversal path opens. One\nof the most outstanding, open questions is about the\nconditions or criteria for the onset of the exchange\ndominated relaxation regime. Crucially, it is unclear\nhow previous understanding gained from observation\nin GdFeCo translates into the recent discovery of\nsingle-pulse switching in the ferrimagnetic Heusler alloy\nMn2RuxGa [35], where the two antiferromagnetically\ncoupled Mn atoms are of the same kind in comparison\nto Gd and FeCo.\nIn this work we present a general theoretical\nframework for the description of single pulse switching\nof ferrimagnets. We provide explicit expressions for\nthe relativistic and exchange relaxation parameters as a\nfunction of microscopic material parameters, including\ntheir dependence on temperature and non-equilibriumsublattice magnetization. This allows us to uncover\nthe criteria for the onset of the exchange-dominated\nrelaxationregimeandswitching. Weverifythevalidityof\nthe model by direct comparison to atomistic spin model\nsimulations of both GdFeCo and Mn 2RuxGa alloys.\nWe shall describe each magnetic atom at site ias a\nclassical spin vector sof a unit length. The magnetic and\nmechanical moments of each atom/element are given by\n\u0016=\u0016ssandS=\u0016ss=\rwhere\u0016sisthemagneticmoment\nand\ris the gyromagnetic ratio. We consider a classical\nHeisenberg spin Hamiltonian[36]:\nH=\u0000X\ni6=jhijiJijsi\u0001sj\u0000X\nidz\ni(sz\ni)2:(1)\nTo model a ferrimagnet, one needs to consider two\nsublattices with different and antiparallel magnetic\nmoments\u0016aand\u0016b, with three different exchange\ncouplingconstants. Twoferromagneticcouplingsforeach\nsublattice coupling with itself ( JaandJb>0) and a\nthirdfortheantiferromagneticinteractionbetweenthem,\nJab<0[37]. The second term in Eq. (1) describes the\ncontribution to the energy of on-site uniaxial anisotropy\nwith easy-axis in z-direction and anisotropy constant, dz\ni.\nThe macroscopic model we propose is derived from a\nmicroscopic spin model, where the equation of motion for\nthe spin dynamics of each atomic spin is the stochastic\nLandau-Lifshitz-Gilbert (LLG) equation [36]:\n@si\n@t=\u0000j\rj\n(1 +\u00152)\u0016i[(si\u0002Hi)\u0000\u0015(si\u0002(si\u0002Hi))]:\n(2)\n\u0015is the local intrinsic atomic damping, the effective field\nHi=\u0010i\u0000@H\n@si, wherethermalfluctuationsarerepresented\nby the stochastic field \u0010i.\nThe non-equilibrium macroscopic magnetization dy-\nnamics of the element-specific angular momentum Sa=\n\u0016ahsai=\r, wherema=hsaiis the first moment of the\nnon-equilibrium distribution function, can be describedarXiv:2201.09067v1 [cond-mat.mtrl-sci] 22 Jan 20222\nelectrons \nlattice \nFIG. 1. Degrees of freedom involved in single-pulse switching;\nelectrons, lattice and element-specific spin systems. Electrons\nact as heat-bath to the spin system with an element-specific\ncoupling strength given by \u000ba(b). Exchange relaxation rate\n\u000bex\u0018\u000ba=ma+\u000bb=mb, wherema(b)are the normalized\nsublattice magnetization magnitude.\nby [19]:\ndSa\ndt=\u000ba\u0016aHa+\u000bex(\u0016aHa\u0000\u0016bHb)(3)\nwherea6=b. The macroscopic relativistic damping\nparameter in Eq. (3) reads [38]\n\u000ba= 2\u0015aL(\u0018a)\n\u0018a: (4)\nHere,L(\u0018)stands for the Langevin function. The\nrelaxation parameter strongly depends on temperature\nand non-equilibrium magnetization state through the\nthermal field \u0018a=\f\u0016aHMFA\na. In the exchange\napproximation, the MFA field acting on sublattice ais:\n\u0016aHMFA\na =zaJaama+zabJabmb: (5)\nHere,zais the number of nearest neighbours (n.n.) of\nspins of type a, andzabis the amount of n.n. of type\nb. The macroscopic damping increases with temperature\nup to a value \u000ba= 2=3at the critical temperature [38].\nThe exchange relaxation parameter \u000bexis given by\n\u000bex=1\n2\u0012\u000ba\nzabma+\u000bb\nzabmb\u0013\n: (6)\nThisexpressionistheextensionofthenon-localexchange\nrelaxation in ferromagnets to local exchange relaxation\nin ferrimagnets. The role of \u000bex,\u000baand\u000bbas the\ncoupling between the sublattices and heat baths is\nvisualized in Fig. 1. In single species ferromagnets,\n0 0.2 0.4 0.6 0.8 1ma00.20.40.60.81mb0.1110\nFIG. 2. Normalized exchange relaxation parameter \u000bex=\u0015\nas function of the sublattice magnetization maandmb, at\nT= 600K. System parameters correspond to GdFeCo.\n\u0015a=\u0015b=\u0015= 0:01. The dotted black line corresponds\nto\u000bex=\u000ba. The white dot represents the starting\nsublattice magnetization ( ma;mb). The closed dashed orange\nline describes a trajectory meeting a no-switching criteria.\nThe open solid white line describes a trajectory meeting a\nswitching criteria.\nsublattices aandbrepresent the same spin lattice,\nhence\u000ba=\u000bb. Therefore, \u000bex=\u000ba=(zma), and\n\u0016aHa\u0000\u0016bHb=\u0016aHexa2\n0\u0001ma, witha0representing\nthe lattice constant. Hence, \u0000non\u0000loc:\nex =\u000bex(\u0016aHa\u0000\n\u0016bHb) =\u000ba(A=M a(T))\u0001ma, whereAis the so-\ncalled micromagnetic exchange stiffness [39]. Ma(T) =\n(\u0016a=\u001da)mais the magnetization density at temperature\nT, where\u001dais the unit cell volume. Non-local exchange\nrelaxation plays a minimal role in the field of ultrafast\nmagnetization dynamics since \u0001ma\u001c1.\nThe non-equilibrium fields ( \u0016aHa= 0at equilibrium)\nare given by\n\u0016aHa=(ma\u0000L(\u0018a))\n\fL0(\u0018a): (7)\nwhere,L0(\u0018) =dL=d\u0018. Notethattheyaredifferenttothe\nMFA fields in Eq. (5), and have been derived previously\n[38, 40]. As the magnetic system approaches thermal\nequilibrium, the non-equilibrium fields can be cast into\nLandau-like expressions [19, 38].\nEquation (3) has been proposed before based on\nsymmetry arguments [19, 23] as a direct generalization\nof the Landau-Lifshitz equation with longitudinal\nrelaxation terms. These models have introduced the\nrelaxation parameters at a purely phenomenological level\nand to some extent their values are arbitrary. Moreover,\nsince they were taken as constant values, most of the\nnon-equilibrium spin physics was not taken into account.\nOur model overcomes these assumptions by providing\nexpressions for the relativistic and exchange relaxation\nparameters as a function of the sublattice specific3\natomic relaxation parameter, \u0015a(b), and normalized\nmagnetization ma(b).\nThis insight is paramount to find the criteria for\nthe onset of a exchange relaxation dominated state.\nIn Fig. 2 we present a diagram of the normalized\nexchange relaxation parameter \u000bex=\u0015for GdFeCo alloy\nparameters as a function of maandmbat a fixed\ntemperature, T= 600K, which corresponds to the\nCurie temperature of the alloy. We observe that the\nexchange relaxation parameter strongly depends on the\nmagnitude of sub-lattice magnetization and its value\nranges over three orders of magnitude, \u000bex=\u0015= 0:1\u0000\n10. Importantly, only when magnetic states reduce\nsignificantly, does the exchange relaxation dominates\nover relativistic relaxation.\nSo far only two classes of ferrimagnets have\nshown single-pulse switching, Gd x(FeCo) 1\u0000xalloys and\nMn2RuxGa. Switching in GdFeCo has been thoroughly\nstudied both experimentally [1–5] and theoretically [3,\n11, 20–23, 31, 32], whereas in Mn 2RuxGa has been only\nrecently demonstrated for a range of Ru concentrations\n(x\u00150:9) [35]. In GdFeCo alloy, switching is\ncharacterised by a fast response of the Fe sublattice and\nslower response of the Gd sublattice to femtosecond laser\nphoto-excitation. This difference roots to their distinct\nmagnetic moment, \u0016Gd= 7:6\u0016Band\u0016Fe= 1:92\u0016B.\nDifferently to GdFeCo, antiferromagnetically coupled\nMn spins in Mn 2RuxGa have similar atomic magnetic\nmoments. We demonstrate the universality of our theory\nby direct comparison of the photo-induced magnetization\nswitching to computer simulations based on atomistic\nspin dynamics (Eq. (2)) for GdFeCo (disordered spin\nstructure) and Mn 2RuxGa (ordered spin structure).\nTheir magnetic properties such as magnetic moments\nand exchange interactions differ substantially. We\nconsider two typical compositions, that show switching,\nGd25(FeCo) 75alloysandMn 2Ru0:86Ga. Itisnoteworthy;\nthat while for GdFeCo atomistic spin models have been\nused thoroughly, in Mn 2Ru0:86Ga similar simulations are\nmissing. We derive the necessary material parameters\nbased on experimental measurements [41–45]. We find\nthat for Mn 2Ru0:86Ga the atomic magnetic moments,\n\u00164a= 2.88\u0016Band\u00164c=4.05\u0016B, and the exchange\nparameters, Ja= 1:28\u00021021J,Jb= 4:0\u00021022J\nandJab=\u00004:85\u00021022J describe the equilibrium\nmagnetization well as function of temperature. Using\nthese parameters the Curie temperature becomes Tc=\n600K and the compensation temperature TM\u0019300K.\nWe note that in the experimental samples, those\ntemperatures are sensitive to the growth conditions as\nwell as material composition. However our temperatures\nare within the reported range of experimentally found\ntemperatures [41–45]. We use the so-called two-\ntemperature model (TTM) to describe the dynamics\nof the electron and phonon systems[46, 47]. For both\nmaterials we use the same parameters in the TTM andthus the electron and phonon temperatures are the same\nfor the same fluence in both materials [45]. The heat-\nbath to which the spins are coupled is represented by the\nelectron system.\nFigures 3 (a) and (b) show excellent agreement\nbetween our macroscopic model and atomistic spin\ndynamics simulations for both alloys for all stages of\nthe magnetization dynamics leading to switching, from\nfast demagnetization, transient ferromagnetic-like state,\nto magnetization recovery. Interestingly, the switching\ndynamics in Mn 2Ru0:86Ga (Fig. 3 (a)) differs to GdFeCo\n(Fig. 3 (b)), both Mn sublattices demagnetize at\nsimilar rate and switch almost simultaneously. Although\ndemagnetization timescales are similar in Mn 2Ru0:86Ga\nand GdFeCo, the recovery of the magnetization in\nMn2Ru0:86Ga is significantly slower. The relaxation of\nthesublatticemagnetization(Eq. 3)canbesplitintotwo\ncontributions, the relativistic relaxation: \u0000r\na=\u000ba\u0016aHa,\nand exchange relaxation \u0000ex=\u000bex(\u0016aHa\u0000\u0016bHb)(see\nFig. 3(c) and (d)). For both ferrimagnetic alloys, \u0000r\na\ndrivesthedynamicsinthefirsthundredsoffemtoseconds,\nuntil sublattice magnetization reduces sufficiently to\nenter the exchange relaxation dominated region \u0000ex>\u0000r\na\n(Fig. 2). In this regime, the exchange relaxation steers\nthe systems towards an intermediate metastable state\ndefined by the condition \u0016aHa=\u0016bHb. Under some\nconditionsthisintermediatestateprecedeswitching. Itis\ncumbersome however to directly analyse Eqs. (7) due to\nitshighlynon-linearcharacter. Thereforeinthefollowing\nwe investigate some limiting scenarios.\nIn the high temperature limit, \u0018a=\f\u0016aHMFA\na!\n0,\u0000r\na= 2\u0015akBTma, which is the so-called thermal\nfluctuation field. The corresponding relaxation time,\n\u001ca=\u0016a=(\r2\u0015akBT), associated to relativistic relaxation,\nhas been discussed before [19, 48]. Similarly, we can\nestimate the high-temperature limit of the exchange\nrelaxation rate\n\u0000ex\n1(ma;mb) =\u0015kBT\nz(ma+mb)(mz\na\u0000mz\nb)\nmamb:(8)\nHigh temperature limits are valid when the temperature\nis larger than the exchange energy acting on the spins.\nOtherwise, intermediate-to-high temperature limit are\nnecessary. This limit adds corrections to previous\nestimations. For instance, \u0016aHa\u0000\u0016bHb= 3kB((T\u0000\nTa\nc)ma+ (T\u0000Tb\nc)mb), whereJ0;a+J0;ab= 3kBTa\nc. The\nexchange relaxation rate is\n\u0000ex\nT(ma;mb) = \u0000ex\n1\u0012\n1\u00001\nTTa\ncma+Tb\ncmb\nma+mb\u0013\n(9)\nFor element-specific critical temperature Ta\nc\u0019Tb\nc, the\ncorrection can be cast as \u0000ex\nT= \u0000ex\n1(T\u0000Ta\nc)=T.Since\nthe relativistic relaxation rate is also modified as \u0000r\na=\n\u0000r\na;1(T\u0000Ta\nc)=T, we can fairly investigate the crossover\nfrom relativistic to exchange dominated regime by4\n0123456FeGd\ntime [ps]−ΓrFeΓrGdΓexexchange relaxationexchange relaxationrelativistic relativistic relativistic −1−0.500.51\n0123\n0123456magnetizationMn4aMn4cΓ[µB/ps]time [ps]−Γr4aΓr4cΓex\n(a)(c)(d)(b)\nFIG. 3. Element-specific magnetization dynamics of single-pulse switching in ferrimagnets using an atomistic spin dynamics\nmodel (symbols) and a macroscopic model (solid lines) for Mn 2Ru0:86Ga (a) and Gd 25FeCo (b). The red area corresponds to\nthe critical values ( mc\na;mc\nb) to enter exchange relaxation regime. The blue area corresponds to the time lapse where relativistic\nrelaxation dominates. Different contributions to the element-specific magnetization relaxation for the single-pulse switching\ndynamics in Mn 2Ru0:86Ga (c) and Gd 25FeCo (d). \u0000r\na(b)stands for the relativistic relaxation rate of sublattice a= 4a(c)\nanda=Fe (d), and b= 4c(c) andb=Gd (d). \u0000exis the exchange relaxation rate, which is the same for both sublattices.\n\u0015a=\u0015b=\u0015= 0:01.\ncomparing \u0000ex\n1and\u0000r\na;1. Two cases of interest exist,\ni) one sublattice is faster than the other, and ii) both\nsublattices demagnetize at the same rate.\nIn the first case, \u001ca\u001c\u001cb, sublattice ademagnetizes\nfaster than sublattice b. This corresponds to GdFeCo\n(Fig. (3)(b)). Soon after the application of a fs\nlaser pulse, ma\u001cmb, and consequently, \u0000ex\u0019\n\u0015kBTmb=(zma)(cf. Eq.(8)). We estimate the conditions\nforthetransitionfromrelativistictoexchange-dominated\nregimes (see Fig. (3)(d)): From \u0000ex>\u0000r\na,ma\u0000r\nbandma\u0014\n1=2z= 0:0833(red colored area in Fig. 3 (b)). A second\ncase,\u001ca\u0019\u001cb, might also arise when demagnetization\ntimes of both sublattices are similar. This is the case\nof Mn 2Ru0:86Ga alloys (Fig. (3)(c)). One can estimate\nthe conditions for the transition \u0000ex= \u0000r\na. This\nhappens for both sublattices when ma;b= (ma(0) +\nmb(0))2=(ma(0)mb(0))=2zab. Assuming a realistic value\nofmb(0)=ma(0) = 0:9, the exchange-dominate regime is\nreached when ma;b= 0:334(red colored area in Fig. 3\n(c)). Notably, this condition only depends on the initial\nvalues of the magnetization, ma(b)(0). These results are\nsimple and general, and one of the main result of the\npresent work. One can interpret the transition from\nrelativistic to exchange-dominated regime as follows:\nas the magnetization of one sublattice decreases, the\nphase space of available states for the spins of the other\nsublattice to switch by exchange of angular momentum\ndramaticallyincreases. Whileswitchingspinviacouplingto an external bath becomes increasingly difficult as the\nmagnetization reduces.\nOnce the system has entered the exchange dominated\nregime the dynamics follows a path where total angular\nmomentum is conserved towards a magnetic state where\n\u0016aHa=\u0016bHb. In the high temperature limit, this\ncondition reduces to mz\na=mz\nb, meaning that the\nexchange relaxation drives the magnetization of both\nsublattices into the same polarity, i.e. a ferromagnetic-\nlike state [2]. From this condition arises, that in order to\nhave a final mz\na<0the following condition is necessary:\nSex\na+Sex\nb<0. HereSex\na(b)standsfortheangularmagnetic\nmoment when the system enters the exchange dominated\nregime. For example, in Mn 2RuxGa, since Mn spins are\nassumed to demagnetize at the same rate, this condition\nreduces to Sex\na+Sex\nb\u0019(Sa(0) +Sb(0)) exp(\u0000t=\u001ca)<\n0. Namely, only for a starting temperature below the\ncompensation temperature, Sa(0)+Sb(0)<0, conditions\nfor switching are fulfilled, in complete agreement to\nexperimental observations [35]. Yet, the exchange\nrelaxation regime needs to be active for a significant time\nin order for the magnetization to switch. We compare\nthe time scales associated to both the relativistic and\nexchange relaxation. Relativistic relaxation rate follows\n\u0000r\na= 2\u0015akB(T\u0000Ta\nc)mz\na, which is strongly reduced by\nthe ultrafast dynamics of mz\na, in only a few hundred of\nfemtoseconds \u0000r\na!0. Whereas \u0000exrather follows the\ndynamics of the temperature T, and therefore decays\nslower than \u0000r\na(Fig. (3)(c) and (d)).5\nThe characteristic time scale of the electron and lattice\ntemperature is described by the TTM and for common\nparameter values in the range of 2 \u00003 ps. As the\ntemperature reduces, the exchange relaxation drives the\nsystem towards (T\u0000Ta\nc)mz\na= (T\u0000Tb\nc)mz\nb. For\nTb\nc< T < Ta\nc, the exchange relaxation drives the\nsystem back into an antiferromagnetic, but switched\nsate. As the magnetization builds up in the opposite\ndirection, \u0000ex, decreases and the relativistic relaxation\ntakes over. Interestingly, our theory predicts that for\nsystems where Ta\nc\u0019Tb\ncswitching would be unlikely,\nwhich has been recently demonstrated in rare-earth free\nsynthetic ferrimagnets [49]. Further, one can accelerate\nthe transition from exchange relaxation dominated to\nthe relativistic regime, and speed up complete switching\nby increasing difference between Ta\ncandTb\nc, an effect\nwhich has been observed by the substitution of Fe by\nCo, namely, GdFe by GdCo [50]. While GdFe recovers\nin tens of picoseconds, GdCo alloys only need of a couple\nof picoseconds.\nTo summarize, in this work we have proposed a general\nmacroscopic theory for the magnetization dynamics\nof ferrimagnetic materials driven by femtosecond laser\nphoto-excitation. Our theory reproduces quantitatively\nall stages of the switching process observed in ex-\nperiments. Notably, we have directly compared our\ntheory to computer simulations using atomistic spin\ndynamics methods for both GdFeCo and Mn 2RuxGa\nalloys. The magnetization dynamics transits from a\nrelativistic relaxation path to an exchange dominated\nregime due to the strong enhancement of the exchange\nrelaxation. We demonstrate that switching occurs when\nthe sublattice magnetization reaches a threshold value.\nThese criteria substitute previous ones and pave the way\nfor the discovery of new class of ferrimagnets showing\nswitching.\nAcknowledgement. The authors thank Ilie Radu and\nJon Gorchon for useful discussions and critical reading\nof the manuscript. We gratefully acknowledge sup-\nport by the Deutsche Forschungsgemeinschaft through\nSFB/TRR 227 \"Ultrafast Spin Dynamics\", Project A08.\n\u0003unai.atxitia@fu-berlin.de.\n[1] C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk,\nA. Tsukamoto, A. Itoh, and T. Rasing, Phys Rev Lett\n99, 047601 (2007).\n[2] I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius,\nH. A. Dürr, T. A. Ostler, J. Barker, R. F. L. Evans,\nR. W. Chantrell, A. Tsukamoto, A. Itoh, A. Kirilyuk,\nT. Rasing, and A. V. Kimel, Nature 472, 205 (2011).\n[3] T. A. Ostler, J. Barker, R. F. L. Evans, R. W. Chantrell,\nU. Atxitia, O. Chubykalo-Fesenko, S. El Moussaoui,\nL. Le Guyader, E. Mengotti, L. J. J. Heyderman,\nF. Nolting, A. Tsukamoto, A. Itoh, D. Afanasiev, B. A.\nIvanov, A. M. Kalashnikova, K. Vahaplar, J. Mentink,A. Kirilyuk, T. Rasing, and A. V. Kimel, Nature\nCommunications 3, 666 (2012).\n[4] L. Le Guyader, S. El Moussaoui, M. Buzzi, R. V.\nChopdekar, L. J. Heyderman, a. Tsukamoto, A. Itoh,\nA. Kirilyuk, T. Rasing, A. V. Kimel, and F. Nolting,\nApplied Physics Letters 101, 022410 (2012).\n[5] C. E. Graves, A. H. Reid, T. Wang, B. Wu,\nS. de Jong, K. Vahaplar, I. Radu, D. P. Bernstein,\nM.Messerschmidt, L.Müller, R.Coffee, M.Bionta, S.W.\nEpp,R.Hartmann,N.Kimmel,G.Hauser,A.Hartmann,\nP. Holl, H. Gorke, J. H. Mentink, A. Tsukamoto,\nA. Fognini, J. J. Turner, W. F. Schlotter, D. Rolles,\nH. Soltau, L. Strüder, Y. Acremann, A. V. Kimel,\nA. Kirilyuk, T. Rasing, J. Stöhr, A. O. Scherz, and H. A.\nDürr, Nature Materials 12, 293 (2013).\n[6] T.-M. M. Liu, T. Wang, A. H. Reid, M. Savoini, X. Wu,\nB. Koene, P. Granitzka, C. E. Graves, D. J. Higley,\nZ. Chen, G. Razinskas, M. Hantschmann, A. Scherz,\nJ. Stoehr, A. Tsukamoto, B. Hecht, A. V. Kimel,\nA. Kirilyuk, T. Rasing, and H. A. Duerr, Nano Letters\n15, 6862 (2015).\n[7] M. S. El Hadri, P. Pirro, C.-H. Lambert, S. Petit-\nWatelot, Y. Quessab, M. Hehn, F. Montaigne, G. Mali-\nnowski, and S. Mangin, Phys. Rev. B 94, 064412 (2016).\n[8] M. L. M. Lalieu, M. J. G. Peeters, S. R. R. Haenen,\nR.Lavrijsen, andB.Koopmans,Phys.Rev.B 96,220411\n(2017).\n[9] J.-Y. Chen, L. He, J.-P. Wang, and M. Li, Physical\nReview Applied 7, 21001 (2017).\n[10] M. L. M. Lalieu, R. Lavrijsen, and B. Koopmans, Nat\nCommun 10(2019), https://doi.org/10.1038/s41467-\n018-08062-4.\n[11] E. Iacocca, T.-M. M. Liu, A. H. Reid, Z. Fu, S. Ruta,\nP. W. Granitzka, E. Jal, S. Bonetti, A. X. Gray,\nC. E. Graves, R. Kukreja, Z. Chen, D. J. Higley,\nT. Chase, L. Le Guyader, K. Hirsch, H. Ohldag, W. F.\nSchlotter, G.L.Dakovski, G.Coslovich, M.C.Hoffmann,\nS. Carron, A. Tsukamoto, A. Kirilyuk, A. V. Kimel,\nT. Rasing, J. Stöhr, R. F. L. Evans, T. Ostler, R. W.\nChantrell, M. A. Hoefer, T. J. Silva, and H. A. Dürr,\nNat Commun 10, 1756 (2019).\n[12] A. V. Kimel and M. Li, Nature Reviews Materials 4, 189\n(2019).\n[13] D. O. Ignatyeva, C. S. Davies, D. A. Sylgacheva,\nA. Tsukamoto, H. Yoshikawa, P. O. Kapralov, A. Kir-\nilyuk, V. I. Belotelov, and A. V. Kimel, Nature\nCommunications 10, 4786 (2019).\n[14] Y. L. W. van Hees, P. van de Meugheuvel, B. Koopmans,\nand R. Lavrijsen, Nature Communications 11, 3835\n(2020).\n[15] C. S. Davies, G. Bonfiglio, K. Rode, J. Besbas,\nC. Banerjee, P. Stamenov, J. M. D. Coey, A. V. Kimel,\nand A. Kirilyuk, Physical Review Research 2, 32044\n(2020).\n[16] Y. Yang, R. B. Wilson, J. Gorchon, C.-H. Lambert,\nS. Salahuddin, and J. Bokor, Science Advances 3,\ne1603117 (2017).\n[17] A. El-Ghazaly, J. Gorchon, R. B. Wilson, A. Pattabi,\nand J. Bokor, Journal of Magnetism and Magnetic\nMaterials 502, 166478 (2020).\n[18] K. Vahaplar, A. Kalashnikova, A. V. Kimel, D. Hinzke,\nU. Nowak, R. Chantrell, A. Tsukamoto, A. Itoh,\nA. Kirilyuk, and T. Rasing, Physical Review Letters\n103, 117201 (2009).6\n[19] J. H.Mentink, J.Hellsvik, D. V.Afanasiev, B.A. Ivanov,\nA. Kirilyuk, A. V. Kimel, O. Eriksson, M. I. Katsnelson,\nand T. Rasing, Physical Review Letters 108, 057202\n(2012).\n[20] J. Barker, U. Atxitia, T. A. Ostler, O. Hovorka, R. W.\nChantrell, O. Chubykalo-Fesenko, and R. W. Chantrell,\nScientific reports 3, 3262 (2013).\n[21] S.Wienholdt, D.Hinzke, K.Carva, P.M.Oppeneer, and\nU. Nowak, Physical Review B 88, 020406 (2013).\n[22] A. J. Schellekens and B. Koopmans, Phys. Rev. B 87,\n020407 (2013).\n[23] V. G. Baryakhtar, V. I. Butrim, and B. A. Ivanov, JETP\nLetters 98, 289 (2013).\n[24] A. Baral and H. C. Schneider, Physical Review B 91,\n100402 (2015).\n[25] V. N. Gridnev, Journal of Physics: Condensed Matter\n28, 476007 (2016).\n[26] A. M. Kalashnikova and V. I. Kozub, Physical Review B\n- Condensed Matter and Materials Physics 93, 1 (2016),\narXiv:arXiv:1506.06585v2.\n[27] G. P. Zhang, T. Latta, Z. Babyak, Y. H. Bai, and T. F.\nGeorge, Modern Physics Letters B 30, 16300052 (2016).\n[28] V. N. Krivoruchko, Physical Review B 94, 54434 (2016).\n[29] S. Gerlach, L. Oroszlany, D. Hinzke, S. Sievering,\nS. Wienholdt, L. Szunyogh, and U. Nowak, Physical\nReview B 95, 224435 (2017).\n[30] V. N. Gridnev, Phys. Rev. B 98, 014427 (2018).\n[31] M. Beens, M. L. M. Lalieu, A. J. M. Deenen, R. A.\nDuine, and B. Koopmans, Physical Review B 100,\n220409 (2019).\n[32] C. Vogler, C. Abert, F. Bruckner, and D. Suess, Physical\nReview B 100, 54401 (2019).\n[33] C. S. Davies, T. Janssen, J. H. Mentink, A. Tsukamoto,\nA. V. Kimel, A. F. G. van der Meer, A. Stupakiewicz,\nand A. Kirilyuk, Phys. Rev. Applied 13, 24064 (2020).\n[34] K. Carva, P. Baláž, and I. Radu (Elsevier, 2017) pp.\n291–463.\n[35] C. Banerjee, N. Teichert, K. E. Siewierska, Z. Gercsi,\nG. Y. P. Atcheson, P. Stamenov, K. Rode, J. M. D. Coey,\nand J. Besbas, Nature Communications 11, 4444 (2020).\n[36] U. Nowak, Classical Spin Models in Handbook of\nMagnetism and Advanced Magnetic Materials (John\nWiley and Sons, Ltd 2007) (2007).\n[37] T. A. Ostler, R. F. L. Evans, R. W. Chantrell,U. Atxitia, O. Chubykalo-Fesenko, I. Radu, R. Abrudan,\nA. Tsukamoto, A. Itoh, A. Kirilyuk, T. Rasing, and\nA. Kimel, Phys. Rev. B 84, 024407.\n[38] U. Atxitia, P. Nieves, and O. Chubykalo-Fesenko,\nPhysical Review B 86, 104414 (2012).\n[39] U. Atxitia, D. Hinzke, O. Chubykalo-Fesenko, U. Nowak,\nH. Kachkachi, O. Mryasov, R. Evans, and R. Chantrell,\nPhysical Review B - Condensed Matter and Materials\nPhysics 82(2010), 10.1103/PhysRevB.82.134440.\n[40] D. A. Garanin, Physical Review B 55, 3050 (1997).\n[41] C. Fowley, K. Rode, Y.-C. Lau, N. Thiyagarajah,\nD.Betto,K.Borisov,G.Atcheson,E.Kampert,Z.Wang,\nY. Yuan, S. Zhou, J. Lindner, P. Stamenov, J. M. D.\nCoey, and A. M. Deac, Phys. Rev. B 98, 220406 (2018).\n[42] D. Betto, N. Thiyagarajah, Y.-C. Lau, C. Piamonteze,\nM.-A. Arrio, P. Stamenov, J. M. D. Coey, and K. Rode,\nPhys. Rev. B 91, 094410 (2015).\n[43] N. Thiyagarajah, Y.-C. Lau, D. Betto, K. Borisov,\nJ. M. D. Coey, P. Stamenov, and K. Rode, Applied\nPhysics Letters 106, 122402 (2015).\n[44] H. Kurt, K. Rode, P. Stamenov, M. Venkatesan, Y.-C.\nLau, E. Fonda, and J. M. D. Coey, Phys. Rev. Lett. 112,\n027201 (2014).\n[45] G. Bonfiglio, K. Rode, G. Y. P. Atcheson, P. Stamenov,\nJ. M. D. Coey, A. V. Kimel, T. Rasing, and A. Kirilyuk,\nJournalofPhysics: CondensedMatter 33,135804(2021).\n[46] M. I. Kaganov, I. M. Lifshitz, and L. V. Tanatarov,\nJETP 173(1957).\n[47] J. K. Chen, D. Y. Tzou, and J. E. Beraun, International\nJournal of Heat and Mass Transfer 49, 307 (2006).\n[48] I.Radu, C.Stamm, A.Eschenlohr, F.Radu, R.Abrudan,\nK. Vahaplar, T. Kachel, N. Pontius, R. Mitzner,\nK. Holldack, A. Föhlisch, T. a. Ostler, J. H. Mentink,\nR. F. L. Evans, R. W. Chantrell, A. Tsukamoto, A. Itoh,\nA. Kirilyuk, a. V. Kimel, and T. Rasing, SPIN 5,\n1550004 (2015).\n[49] J.-W. Liao, P. Vallobra, L. O’Brien, U. Atxitia,\nV. Raposo, D. Petit, T. Vemulkar, G. Malinowski,\nM. Hehn, E. Martínez, S. Mangin, and R. Cowburn,\nAdvanced Science 6(2019), 10.1002/advs.201901876.\n[50] A. Ceballos, A. Pattabi, A. El-Ghazaly, S. Ruta, C. P.\nSimon, R. F. L. Evans, T. Ostler, R. W. Chantrell,\nE. Kennedy, M. Scott, J. Bokor, and F. Hellman,\nPhysical Review B 103, 24438 (2021)." }, { "title": "2303.15191v4.Strain_effects_on_magnetic_compensation_and_spin_reorientation_transition_of_Co_Gd_synthetic_ferrimagnets.pdf", "content": "Strain effects on magnetic compensation and spin reorientation transition of Co/Gd\nsynthetic ferrimagnets\nGiovanni Masciocchi,1, 2, a)Thomas J. Kools,3Pingzhi Li,3Adrien A. D. Petrillo,3Bert Koopmans,3Reinoud\nLavrijsen,3Andreas Kehlberger,2and Mathias Kläui1\n1)Institute of Physics, Johannes Gutenberg University Mainz, Staudingerweg 7, 55128 Mainz,\nGermany\n2)Sensitec GmbH, Walter-Hallstein-Straße 24, 55130 Mainz, Germany\n3)Department of Applied Physics, Eindhoven University of Technology, Eindhoven, 5612 AZ,\nNetherlands\n(Dated: 3 April 2023)\nSynthetic ferrimagnets are an attractive materials class for spintronics as they provide access to all-optical switching\nof magnetization and, at the same time, allow for ultrafast domain wall motion at angular momentum compensation.\nIn this work, we systematically study the effects of strain on the perpendicular magnetic anisotropy and magnetization\ncompensation of Co/Gd and Co/Gd/Co/Gd synthetic ferrimagnets. Firstly, the spin reorientation transition of a bilayer\nsystem is investigated in wedge type samples, where we report an increase in the perpendicular magnetic anisotropy\nin the presence of in-plane strain. Using a model for magnetostatics and spin reorientation transition in this type of\nsystem, we confirm that the observed changes in anisotropy field are mainly due to the Co magnetoelastic anisotropy.\nSecondly, the magnetization compensation of a quadlayer is studied. We find that magnetization compensation of this\nsynthetic ferrimagnetic system is not altered by external strain. This confirms the resilience of this material system\nagainst strain that may be induced during the integration process, making Co/Gd ferrimagnets suitable candidates for\nspintronics applications.\nI. INTRODUCTION\nRecent advances in spintronics have opened new possi-\nbilities for electronic applications beyond the CMOS stan-\ndard. New concepts of high density and ultrafast non-volatile\ndata storage have been proposed in magnetic memories1,2.\nThroughout the years, magnetic memories have evolved3,4\nexploiting different geometries5and new material platforms\nsuch as ferrimagnets6have been used to improve storage\ndensity7, reading and writing speed8and energy efficiency9,10.\nAt the same time, single-pulse optical-switching (AOS) of\nmagnetization has reduced the switching speed of the mag-\nnetization below ps timescale11–14. This bears promise for a\nnew generation of ultrafast data buffering, in a single chip that\nintegrates photonics with spintronics15–19.\nFerrimagnets are a class of magnets with unbalanced\nantiparallel-aligned sublattice moments. The compensation\nof the two inequivalent sublattices, combines the advantages\nof both antiferromagnets (antiparallel alignment of magnetic\nmoments) and ferromagnets (finite Zeeman coupling and spin\npolarization)16,20. Moreover, the drastic contrast between the\ntwo sublattices in non-adiabatic dynamics, could potentially\naccommodate AOS by a femtosecond laser pulse12,16. Single-\npulse AOS is typically observed in rare earth–transition metal\n(RE–TM) ferrimagnetic alloys like GdFeCo20or in multilayer\nsynthetic ferrimagnet, such as Co/Gd and [Co/Tb] n21,22. In\nparticular, the one based on multilayer of Co/Gd is a good can-\ndidate for integrated opto-spintronics devices as it shows AOS\n- without the constrains on the composition as imposed by al-\nloy system23,24- and at the same time exhibits magnetic and\nangular momentum compensation, allowing ultrafast domain\na)Electronic mail: gmascioc@uni-mainz.dewall motion25,26. For instance, the integration of Co/Gd syn-\nthetic ferrimagnets in an optically switchable magnetic tunnel\njunction has been recently reported27.\nWhen it comes to technological implementation, strain\ninduced effects must be considered, which could be in-\ncurred from processing steps such as packaging and layer\ndeposition28. Intrinsic stresses and strain could affect the\nmagnetic anisotropy via changes to the spin-orbit coupling\n(SOC)29or to the magnetization compensation of ferrimag-\nnets especially in RE–TM alloys30,31. However, in spite of\nbeing omnipresent in applications32–34, the effect of strain has\nnot yet been explored in these materials. In this work, we\npresent a systematic study of the effects of strain on Co/Gd\nsynthetic ferrimagnets. By the application of external strain,\nusing substrate bending, we investigate the impact of strain on\nthe perpendicular magnetic anisotropy (PMA) and the mag-\nnetization compensation of [Co/Gd] and [Co/Gd] 2multilay-\ners, respectively. Using wedge samples in a bilayer system of\nCo/Gd and polar magneto-optic Kerr effect (pMOKE) mea-\nsurements, we confirm that the PMA is increased by in-plane\ntensile strain and a negative magnetostriction is reported. By\nincluding the contribution of the strain-anisotropy for this sys-\ntem in a model for the magnetostatics, we show that the ef-\nfects of strain on the magnetization are mainly due to the\nmodification of the spin-orbit coupling within the magnetic\nlayer and at the the Pt/Co interface that increases the mag-\nnetic anisotropy via magnetoelastic coupling. Additionally,\nwe find that the magnetization compensation point is not af-\nfected significantly by strain, as the magnetoelastic coupling\naffects the anisotropy rather than the magnetization of the two\nsublattices. Our study explores the mechanisms that underlie\nthe influence of strain on the magnetic anisotropy of Co/Gd\nferrimagnets and contributes to a better understanding of the\nmagnetoelastic effects of ferrimagnetic multilayers. These re-\nsults could be employed for the optimization and developmentarXiv:2303.15191v4 [cond-mat.mtrl-sci] 31 Mar 20232\nof spintronics devices, as well as for potential applications in\nfields such as magnetic memory and sensing.\nII. METHODS AND SAMPLE FABRICATION\nThe samples were grown on a 1.5 µm thick, thermally oxi-\ndized SiOx on top of a 625 µm thick Si substrate by DC mag-\nnetron sputtering in a chamber with a typical base pressure\nof 5×10−9mBar. To obtain a variable thickness (wedge)\nalong the sample surface, a shutter in the close proximity of\nthe sample is gradually closed during deposition. This al-\nlows to study the compensation and spin reorientation tran-\nsition (SRT) within a single sample. Two types of sam-\nples are realized. Firstly, a bilayer of Ta(4 nm)/Pt(4)/ Co(0-\n2)/Gd(t Gd)/TaN(4) with a constant Gd layer on top of a Co\nwedge is considered to study the SRT. In addition, a quadlayer\nof Ta(4)/Pt(4)/Co(0.6)/Gd(0-2)/Co(0.6)/Gd(1.5)/TaN(4), this\ntime with a Gd wedge, is grown to study the magnetization\ncompensation.\nThe magnetic properties of these wedge samples were in-\nvestigated by pMOKE, where we are only sensitive to the\nout-of-plane (OOP) component of the Co magnetization at a\nwavelength of 658 nm. According to Fig. 1 (a), the surface\nof the sample is scanned along the y-direction using a focused\nlaser spot with a spot-size of /similarequal250µm diameter. Accord-\ningly, the local magnetic properties and hysteresis loops can\nbe measured as a function of layer thickness, with a negligi-\nble thickness gradient <0.025 nm within the used laser spot.\nAll the measurements are performed at room temperature. To\napply in-plane tensile strain to our multilayer, the substrate\nis mechanically bent using a three-point method35. A square\nsample of 1 by 1 cm is vertically constrained on two sides\nand pushed uniformly from below by a cylinder that has off-\ncentered rotation axis. The device generates a tensile strain in\nthe plane of the sample when the cylinder is rotated. As pre-\nviously reported, the tensile strain is uniaxial along xand uni-\nform in the measured area of the sample. The in-plane strain\nmagnitude is 0 .1% and has been measured with a strain gauge\n(RS PRO). More details about the strain generating device can\nbe found in section S2 of the supplementary information.\nIII. RESULTS AND DISCUSSION\nA. Spin reorientation transition in Co/Gd bilayer\nThe use of magnetic materials for high density data storage\nrequires magnetic systems that are OOP magnetized36,37. In\nthin films, an OOP magnetic easy axis can be obtained by\nmagnetocrystalline anisotropy induced at the interface with\nheavy metal38,39. In addition to that, strain has been shown\nto affect the magnetic easy axis direction in systems with\nPMA40. To understand the effect of external strain on Co/Gd\nsystems with PMA, we investigate bilayer samples consist-\ning of Ta(4 nm)/Pt(4)/ Co(0-2)/Gd( tGd)/TaN(4). Specifically,\nthe Co thickness is varied between 0 and 2 nm over a few\nmm along the ydirection, whereas tGdis constant (as in Fig.1 (a)). In this system, the balance between the interfacial\nanisotropy energy (magnetocrystalline anisotropy energy at\nthe Pt/Co interface) and demagnetization energy determines\nthe effective magnetic anisotropy. In such system, the demag-\nnetization energy increases with the thickness of the Co mag-\nnetic layer, and consequently, the magnetization will go from\nout-of-plane (OOP) to in-plane (IP). To probe the magnetiza-\ntion of our wedge sample, we record hysteresis loops from\nthe pMOKE signal. We repeat the measurement moving the\nlaser spot along the wedge in the y direction. Firstly, a sample\nwhere t Gd=0 is considered. This measurement can be seen in\nFigs. 1 (b) and (c). Fig. 1 (b) reports the magnetic response of\nthe Ta(4 nm)/Pt(4)/Co(0-2)/TaN(4) sample to an OOP mag-\nnetic film for different t Co. The effective anisotropy Ke f fwas\nestimated38recording hysteresis loops with magnetic field ap-\nplied OOP and IP and the corresponding anisotropy energy\nper unit area is Ks=1.7 mJ/m2. For tCo=1.35 nm the square-\nshaped loop indicates PMA with Ke f f= 1.5(2)×105J/m3. A\nvalue of MCo=1.3 MA/m was used in the calculation. As the\nthickness of Co is increased (moving the laser spot along the\nwedge direction - y) the remanence and squareness of the hys-\nteresis loop decreases together with the PMA of the system.\nFortCo=2.00 nm, the sample is IP magnetized and Ke f f=-\n0.8(2)×105J/m3is negative. The OOP to IP transition occurs\nattCo=1.85(2)nm in this system.\nTo investigate the effects of externally applied in-plane\nstrain, we repeat the measurement while the sample is me-\nchanically bent. The magnetization is coupled to the ex-\nternal strain and can be described by the expression for the\nanisotropy energy35:\nKME=−3\n2λsYε, (1)\nwhere λsis the saturation magnetostriction, Yis the\nYoung’s modulus and εis the strain. If the strain in the film\nis non-zero, the magneto-elastic coupling of Co contributes\nin principle to the effective anisotropy. Accordingly, the total\nanisotropy Ke f fof the magnetic stack is expected to change in\nthe presence of external strain. Fig. 1 (c) shows the OOP hys-\nteresis loops of Ta(4 nm)/Pt(4)/Co(1.85)/TaN(4) sample be-\nfore (blue) and after (red) the application of εxx=0.1%. We\nobserve that the anisotropy field is decreased after the appli-\ncation of in-plane strain. This happens because, in this sys-\ntem, the strain-induced magnetoelastic anisotropy KME=0.02\nmJ/m2is positive, as we expect from a material with negative\nmagnetostriction like Co40,41. More details about the calcu-\nlations of magnetoelastic anisotropy can be found in section\nS2 of the supplementary information. Accordingly, the PMA\nis increased by the applied strain, i.e. the system is expected\nto be OOP magnetized for thicker Co if compared to samples\nwithout strain.\nAfter this preliminary study on Pt/Co systems, we focused\nour attention on the magnetostriction of Co/Gd multilayers.\nIn Co-Gd alloys the magnetostriction has been reported to be\nstrongly dependent on the composition29,42due to the struc-\ntural modification occurring with different atomic content. In\ncontrast to this case, the effects of magnetostriction of a mul-3\nFIG. 1: (a) Sample sketch, red arrow indicates the direction of the applied strain. (b) Out of plane hysteresis loops of a Pt/Co/TaN stack for different Co\nthicknesses. (c) OOP hysteresis loops of Pt/Co(1.85 nm)/TaN before (blue) and after (red) application of 0 .1% in-plane strain. (d) MOKE intensity scan at\nremanence (no applied field) of Pt/Co/Gd/TaN films along the Co wedge.\ntilayer, are expected to be dependent on the magnetoelastic\ncoupling of the individual layers43.\nTo study the magnetostriction of a Co/Gd multilayer, a con-\nstant layer of Gd on top of the Co wedge is added. To perform\nthickness dependent studies, a thickness tGd=1 nm and 3 nm\nis considered. In the bilayer system, the magnetization in the\nGd layers is mainly induced at the interface with the Co layer,\nand couples anti-parallel the Co magnetization21. Accord-\ningly, tCorequired to reach SRT is expected to change with in-\ncreasing tGd44. To compare the SRT of Ta(4 nm)/Pt(4)/Co(0-\n2)/Gd( tGd)/TaN(4) samples with different tGdwe performed\nremanent intensity scan along our Co wedge, in addition\nto hysteresis loop measurements. After the sample is sat-\nurated with an OOP magnetic field of 1T, we determine\nthe thickness-dependent remanence from the pMOKE signal\nwithout external magnetic field. The remanent intensity scans\nare reported in Fig. 1 (d). As the pMOKE signal is mainly\nsensitive to the OOP component of Co magnetization, the nor-\nmalized remanent intensity will drop to zero at the SRT, when\nthe magnetization rotates IP. The SRT can be observed in Fig.\n1 (d) in samples with different thicknesses of Gd before and\nafter the application of strain. As previously reported44the\ncritical thickness tCo=tcat which SRT occurs, changes sig-\nnificantly in the presence of a Gd layer. For all the considered\nsamples, the in-plane strain shifts the OOP to IP transition\ntowards larger Co thickness. This suggests that the effective\nmagnetostriction of the Co/Gd bilayer is negative and its value\nλs=−10(5)×10−6is not significantly altered by the pres-ence of the Gd layer.\nTo obtain a quantitative understanding of the shape of the\nspin reorientation boundary, we employ an analytical model44\ndescribing the magnetostatic free energy of the anisotropy,\nwhich is zero at the SRT boundary. The first constituent ener-\ngies of the model are the demagnetization energies of the Co\nlayer\nEd,Co=1\n2µ0/integraldisplayy\n0M2\nCodq=1\n2µ0M2\nCoy (2)\nand of the Gd layer\nEd,Gd=1\n2µ0/integraldisplayx\n0M2\nGdexp(−2q/λGd)dq=\n1\n4µ0M2\nGdλGd/parenleftbigg\n1−exp/parenleftbigg−2x\nλGd/parenrightbigg/parenrightbigg (3)\nwhere λGdis the characteristic decay length of the Gd mag-\nnetization, which is induced at the Co/Gd interface, MCois the\nmagnetization of the Co layer, MGdis the effective Gd mag-\nnetization at the interface between Co and Gd and xandy\nare, respectively, the Gd and Co thicknesses in the diagram of\nFig.2 (a). The plot axes in Fig.2 (a) have been inverted for a\nbetter comparison with the other figures. The magnetocrys-\ntalline anisotropy is included with the term4\nEK=Ks−∆K/parenleftbigg\n1−exp/parenleftbigg−2x\nλK/parenrightbigg/parenrightbigg\n, (4)\nand it is also considered to decay with a characteristic de-\ncay length λKand magnitude ∆K. The second term in Eq.\n4 phenomenologically addressed the experimentally observed\ndecay in the effective anisotropy, which may be caused by\nsputter induced disordering of the Co45. Using a numeri-\ncal fit to the experimentally determined SRT, the parame-\ntersλK,λGdand∆Kfor our Co/Gd bilayer are determined.\nAll the other parameters were either experimentally measured\nor taken from literature and are reported in Table S.1 , sec-\ntion S1 of the supplementary information. In addition to the\nanisotropy term, and additional energy term Emixis included\nin the model. Emixtakes into account the mixing at the mag-\nnetic layer interfaces where the local net magnetization is\nzero. More details about the expression for this term and the\ndetermination of the fitting parameters can be found in the\nsupplementary information and in the work of Kools et al.44.\nIn this model, the expression of the total free energy density\nper unit area is, considering all the terms mentioned so far:\nEtot=−EK−Emix+Ed,Co+Ed,Gd. (5)\nThe magnetocrystalline anisotropy energy per unit area Ks,\ndue to the Pt/Co interface is assumed constant.\nEq. 5, describing the total energy of a\nTa(4nm)/Pt(4)/Co( tCo)/Gd( tGd)/TaN(4) sample, can be\nsolved for y (t Co) by imposing Etot=0 (spin reorientation\ntransition). The solution for the SRT obtained with the model\ndescribed above is reported in Fig. 2 (a) with a blue solid line\nin a phase diagram where tGd(x) and tCo(y) are continuously\nvaried from 0 to 3 nm and from 0 to 2 nm, respectively.\nTogether with the calculations, the SRT measured experimen-\ntally without externally applied strain is reported with blue\ndiamonds in Fig. 2 (a). The experimental data, follow well\nthe general trend of the calculations. Discrepancies between\nmodel and experimental values for tGd=0, might be due to\nadditional mixing between the layers.\nTo include the effects of strain, a magnetoelastic anisotropy\nKMEis added to Eq. 5 that becomes\nEtot=−EK−Emix−KME+Ed,Co+Ed,Gd. (6)\nIn our case KME=0.02 mJ/m2corresponds to the value of\nmagnetoelastic anisotropy induced with 0 .1% externally ap-\nplied in-plane strain in our experiments. As showed in Fig.\n1 (d), we do not observe significant changes to KMEwith in-\ncreasing tGd. Again considering the SRT-boundary to be at\nEtot=0, the solution of Eq. 6 (that includes the magnetoe-\nlastic term) is reported in Fig. 2 (a) with an orange solid line.\nAs expected from a material with negative magnetostriction,\nKMEsums to Ksand the PMA is enhanced by in-plane strain.\nThe SRT calculated including KMEto Eq. 6 is consequently\nshifted to larger values of tCo. This trend is in agreement with\nFIG. 2: (a) 2D phase diagram of the SRT of the a\nTa(4nm)/Pt(4)/Co( tCo)/Gd(t Gd)/TaN(4) stack as a function of tGd(x) and tCo\n(y). The axes have been inverted for a better comparison with other figures.\nBlue diamonds and red squares correspond to the experimental data,\nreported without and with strain applied, respectively. The solid lines\nindicate the calculated values using the model for the magnetostatics and Eq.\n6. A magnetoelastic anisotropy KME=0 and 0.02 mJ/m2is considered,\nrespectively, for the blue and orange curve. (b) Spin reorientation transition\nof a Ta(4)/Pt(4)/Co( tCo)/Gd(t Gd)/TaN(4) sample calculated for values of\ntGd=0, 1 and 3 nm and plotted as a function of tCo. The SRT is represented\nhere by a step function. Solid and dashed lines consider KME=0 and 0.02\nmJ/m2, respectively.\nthe experimentally determined SRT when and external strain\nεxx=0.1% is applied (orange squares in Fig.2 (a)).\nAnother way to visualize the SRT is solving Eq. 6 for fixed\nvalues of tGdand obtaining the critical thickness of tCosuch\nthatEtot=0. Then, the SRT can be represented as a step\nfunction in the diagram of Fig. 2 (b), analogue to the MOKE\nremanence scan shown in Fig. 1 (d). The values of Gd thick-\nnesses considered are tGd=0, 1 and 3 nm and are plotted in\nFig. 2 (b) with solid lines in black, blue and orange, in order.\nSolid lines consider KME=0 mJ/m2. Dashed lines consider\ninstead KME=0.02 mJ/m2in Fig. 2 (b). The information\ncontained here can be correlated to the experimental rema-\nnent intensity scan in Fig. 1 (d). Comparing Fig. 2 (b) with5\nFIG. 3: (a) Layerstack consisting of a Co/Gd quadlayer used to obtain magnetization compensation. (b) Coercivity and (c) remanent pMOKE intensity scan as\na function of tGd. Measurements before (blue) and after (orange) application of in-plane strain are reported. (d) Hysteresis loops in the Co dominated and (e)\nGd dominated state. Both curves with (orange) and without (blue) in-plane strain applied are shown.\nFig. 1 (d), a similar behavior can be observed. Firstly we can\nnote that the model predicts the SRT to shift when the thick-\nness of the Gd layer is tGd>0. Secondly, we observe a similar\nshift of the SRT point in Fig. 2 (b) and Fig. 1 (d) due to the\neffect of magnetoelastic anisotropy and of the external strain,\nrespectively. As we expect from a material with negative mag-\nnetostriction, Ksadds to KME, therefore the PMA is increased\nand the Co/Gd bilayer stays OOP magnetized for thicker Co\n(corresponding to larger Ed,Co). We confirm that the major ef-\nfect of strain on the Ta(4 nm)/Pt(4)/ Co(0-2)/Gd(t Gd)/TaN(4)\nsample is the alteration of the PMA. Moreover, the estimated\neffective magnetostriction of the stack - λs=−10(5)×10−6\n- is not significantly altered by the presence of the Gd layer in\nthe thickness range considered.\nIn this section, we examined the impact of in-plane strain\non the effective PMA of a Co/Gd ferrimagnetic bilayer. Our\nresults suggest negative magnetostriction of the stack for the\ninvestigated thickness values. We employ a recent model for\nthe magnetostatics of these type of systems, where we include\nthe effects of strain purely as magnetoelastic anisotropy. Our\nexperimental findings are in good agreement with the predic-\ntions made by this model, providing deeper understanding of\nthe response of this material platform to external strain.\nB. Magnetization compensation in quadlayer systems\nIn ferrimagnets, magnetization compensation can be\nachieved. This occurs when the net magnetization /vectorMtot=\n/vectorMGd+/vectorMCovanishes because the magnetization, coming from\nthe two sub-lattices, is equal in magnitude and opposite in\nsign.\nIn recent studies, changes to the saturation magnetization\nin the presence of strain were reported in epitaxial films31\nand rare earth free ferrimagnets30. To study the effectsof strain on magnetization compensation of synthetic fer-\nrimagnets, we consider a quadlayer sample44consisting of\nTa(4 nm)/Pt(4)/Co(0.6)/Gd(0-2)/Co(0.6)/Gd(1.5)/TaN(4) as\nschematically drawn in Fig. 3 (a). In this case, the thickness of\nthe bottom Gd layer is varied between 0 and 2 nm over a few\nmm, whereas all the other layers have constant thickness. The\nreason for this choice is that compared to the Co/Gd bilayer,\nthe magnetic volume of the Co is doubled while the number\nof Co/Gd interfaces where magnetization is induced in the Gd\nthrough direct exchange with the Co, is tripled. In this way\nmagnetization compensation can be more readily achieved.\nThe growing thickness of Gd, increases the contribution of\n/vectorMGdto/vectorMtot. For this reason, some areas of the wedge sam-\nple will be Co-dominated (for tGdtcomp) with /vectorMtot=0 attGd=tcomp.\nHere, tcomp is the thickness where magnetization compensa-\ntion is obtained. At magnetization compensation two effects\nare expected: a divergence of the coercivity and a sign change\nin the remanent pMOKE signal (Kerr rotation, normalized to\nits value in absence of Gd). The measurements for coercivity\nand intensity are reported in Figs. 3 (b) and (c), respectively.\nThe coercivity data were extracted from hysteresis loops mea-\nsured across the wedge direction (along y). The reason for\nthe sign change in the pMOKE signal, is the alignment of the\nGd magnetization along the field direction, in the Gd domi-\nnated regime. We report magnetization compensation in this\nquad-layer for tGd=1.25 nm.\nIn a similar fashion to what we have done investigating the\nPMA in the bilayer system, we repeat the experiment in the\npresence of εxx=0.1% in-plane strain. The results are reported\nin orange in Fig. 3 (b) and (c). Remarkably, the compensation\npoint of the Co/Gd quadlayer is unchanged by the application\nof this externally applied strain.\nFigs. 3 (d) and (e) contain OOP hysteresis loops\nof Ta(4 nm)/Pt(4)/Co(0.6)/Gd( tGd)/Co(0.6)/Gd(1.5)/TaN(4)6\nsamples for tGd=1.15 nm and tGd=1.35 nm, respectively,\nand further show the effects of magnetization compensation.\nThe sample is in this case OOP magnetized. As the thickness\nof Gd is increased, the magnetization of the sample goes from\nCo dominated (Fig. 3 (d)) to Gd dominated (Fig. 3 (e)). The\ninversion of hysteresis loops happens because for tGd>1.25\nnmthe Co-magnetization aligns antiparallel to the field, lead-\ning to the change in sign of the pMOKE signal. When the\nmeasurement is repeated in the presence of εxx=0.1% strain\n(orange line), no significant changes to the remanent intensity\nor coercivity are reported, if compared to the unstrained case\n(blue line). This suggests that magnetization compensation\ncan be achieved in these multilayer systems in the presence of\nexternal strain and, most importantly, that the magnetization\ncompensation point is unaffected.\nTo explain this, we can consider earlier studies about\nmagnetostatics of these types of systems. As previously\nreported25,44, magnetization compensation is due to the bal-\nance in Co magnetization and the Gd magnetization, induced\nin the Gd at the Co/Gd interfaces. In-plane strain in multilayer\nsamples with PMA modifies spin orbit coupling within one\nlayer46, thus altering the magnetocrystalline anisotropy en-\nergy of the system47. On the other hand the total magnetic mo-\nment per unit area /vectorMtotin synthetic ferrimagnets is obtained\nby integrating the magnetization of the Co and Gd sublattices\nover the respective layer thicknesses. Accordingly, in a mul-\ntilayer in-plane strain is not affecting the induced magnetic\nmoment from the Co onto the Gd, thus not altering magneti-\nzation compensation.\nIV. CONCLUSIONS\nThis work reveals the effect that external strain has on PMA\nand magnetization compensation of Co/Gd systems at room\ntemperature. Growing wedge samples, where the thickness of\none of the magnetic layers was varied, has allowed us to deter-\nmine thickness dependent transition in the magnetostatics of\nthis multilayer system. Deliberate in-plane strain was applied\nto the sample. In a bilayer Pt/Co/Gd system, we experimen-\ntally show that a sizable magnetoelastic coupling changes the\nSRT in the presence of strain. The contribution of the strain-\nanisotropy for this system has been included in a model for the\nmagnetostatics, describing the experimental observations well\nif an effective negative magnetostriction is considered. In a\nPt/Co/Gd/Co/Gd quadlayer we obtain magnetization compen-\nsation of the two sub-lattices by varying the thickness of the\nbottom Gd layer. Here, we find that the application of in-plane\nstrain does not affect the magnetization compensation. The\ninduced magnetic moment from the Co onto the Gd, being an\ninterface effect in a multilayer system, is not altered by such\nmechanical deformation. To conclude, this work provides a\nbroad understanding of the magnetoelastic properties of these\nmultilayer systems. As PMA and magnetic compensation are\nmaintained in the presence of externally applied strain, this\nmaterial system is a good candidate for technological imple-\nmentation of ferrimagnets.SUPPLEMENTARY MATERIAL\nSee supplementary material for magnetostatics model for\nthe spin reorientation transition and for more details about the\nsetup used for application of strain.\nACKNOWLEDGMENTS\nThis project has received funding from the European\nUnion’s Horizon 2020 research and innovation program un-\nder the Marie Skłodowska-Curie grant agreement No 860060\n“Magnetism and the effect of Electric Field” (MagnEFi), the\nDeutsche Forschungsgemeinschaft (DFG, German Research\nFoundation) - TRR 173 - 268565370 (project A01 and B02)\nand the Austrian Research Promotion Agency (FFG). The au-\nthors acknowledge support by the Max-Planck Graduate Cen-\ntre with Johannes Gutenberg University.\nAUTHOR DECLARATIONS\nConflict of interest\nThe authors have no conflicts to disclose.\nDATA SHARING POLICY\nThe data that support the findings of this study are available\nfrom the corresponding author upon reasonable reques.\n1T. Endoh, H. Honjo, K. Nishioka, and S. Ikeda, “Recent progresses in\nSTT-MRAM and SOT-MRAM for next generation MRAM,” in 2020 IEEE\nSymposium on VLSI Technology (IEEE, 2020) pp. 1–2.\n2S. S. Parkin, M. Hayashi, and L. Thomas, “Magnetic domain-wall race-\ntrack memory,” Science 320, 190–194 (2008).\n3S. Tehrani, “Status and outlook of MRAM memory technology,” in 2006\nInternational Electron Devices Meeting (IEEE, 2006) pp. 1–4.\n4K. Garello, F. Yasin, and G. S. Kar, “Spin-orbit torque MRAM for ul-\ntrafast embedded memories: From fundamentals to large scale technology\nintegration,” in 2019 IEEE 11th International Memory Workshop (IMW)\n(IEEE, 2019) pp. 1–4.\n5K. Gu, Y . Guan, B. K. Hazra, H. Deniz, A. Migliorini, W. Zhang, and\nS. S. Parkin, “Three-dimensional racetrack memory devices designed from\nfreestanding magnetic heterostructures,” Nature Nanotechnology 17, 1065–\n1071 (2022).\n6S.-H. Yang, K.-S. Ryu, and S. Parkin, “Domain-wall velocities of up to 750\nm s- 1 driven by exchange-coupling torque in synthetic antiferromagnets,”\nNature nanotechnology 10, 221–226 (2015).\n7R. Tomasello, V . Puliafito, E. Martinez, A. Manchon, M. Ricci, M. Car-\npentieri, and G. Finocchio, “Performance of synthetic antiferromagnetic\nracetrack memory: domain wall versus skyrmion,” Journal of Physics D:\nApplied Physics 50, 325302 (2017).\n8S.-H. Yang, C. Garg, T. Phung, C. Rettner, and B. Hughes, “Spin-orbit\ntorque driven one-bit magnetic racetrack devices-memory and neuromor-\nphic applications,” in 2019 International Symposium on VLSI Technology,\nSystems and Application (VLSI-TSA) (IEEE, 2019) pp. 1–2.\n9Q. Shao, Z. Wang, and J. J. Yang, “Efficient AI with MRAM,” Nature\nElectronics 5, 67–68 (2022).\n10S. Parkin and S.-H. Yang, “Memory on the racetrack,” Nature nanotechnol-\nogy10, 195–198 (2015).7\n11I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius, H. Dürr, T. Ostler,\nJ. Barker, R. Evans, R. Chantrell, et al. , “Transient ferromagnetic-like state\nmediating ultrafast reversal of antiferromagnetically coupled spins,” Nature\n472, 205–208 (2011).\n12T. Ostler, J. Barker, R. Evans, R. Chantrell, U. Atxitia, O. Chubykalo-\nFesenko, S. El Moussaoui, L. Le Guyader, E. Mengotti, L. Heyderman,\net al. , “Ultrafast heating as a sufficient stimulus for magnetization reversal\nin a ferrimagnet,” Nature communications 3, 1–6 (2012).\n13A. V . Kimel and M. Li, “Writing magnetic memory with ultrashort light\npulses,” Nature Reviews Materials 4, 189–200 (2019).\n14P. Zhang, T.-F. Chung, Q. Li, S. Wang, Q. Wang, W. L. Huey, S. Yang, J. E.\nGoldberger, J. Yao, and X. Zhang, “All-optical switching of magnetization\nin atomically thin CrI 3,” Nature materials 21, 1373–1378 (2022).\n15E. K. Sobolewska, J. Pelloux-Prayer, H. Becker, G. Li, C. S. Davies,\nC. Krückel, L. A. Félix, A. Olivier, R. C. Sousa, I.-L. Prejbeanu, et al. ,\n“Integration platform for optical switching of magnetic elements,” in Ac-\ntive Photonic Platforms XII , V ol. 11461 (SPIE, 2020) pp. 54–72.\n16S. K. Kim, G. S. Beach, K.-J. Lee, T. Ono, T. Rasing, and H. Yang, “Ferri-\nmagnetic spintronics,” Nature Materials 21, 24–34 (2022).\n17L. Avilés-Félix, L. Álvaro-Gómez, G. Li, C. Davies, A. Olivier, M. Rubio-\nRoy, S. Auffret, A. Kirilyuk, A. Kimel, T. Rasing, et al. , “Integration of\nTb/Co multilayers within optically switchable perpendicular magnetic tun-\nnel junctions,” Aip Advances 9, 125328 (2019).\n18M. L. Lalieu, R. Lavrijsen, and B. Koopmans, “Integrating all-optical\nswitching with spintronics,” Nature communications 10, 110 (2019).\n19H. Becker, C. J. Krückel, D. Van Thourhout, and M. J. Heck, “Out-of-\nplane focusing grating couplers for silicon photonics integration with op-\ntical MRAM technology,” IEEE Journal of Selected Topics in Quantum\nElectronics 26, 1–8 (2019).\n20K.-J. Kim, S. K. Kim, Y . Hirata, S.-H. Oh, T. Tono, D.-H. Kim, T. Okuno,\nW. S. Ham, S. Kim, G. Go, et al. , “Fast domain wall motion in the vicin-\nity of the angular momentum compensation temperature of ferrimagnets,”\nNature materials 16, 1187–1192 (2017).\n21M. Lalieu, M. Peeters, S. Haenen, R. Lavrijsen, and B. Koopmans, “De-\nterministic all-optical switching of synthetic ferrimagnets using single fem-\ntosecond laser pulses,” Physical review B 96, 220411 (2017).\n22L. Avilés-Félix, A. Olivier, G. Li, C. S. Davies, L. Álvaro-Gómez,\nM. Rubio-Roy, S. Auffret, A. Kirilyuk, A. Kimel, T. Rasing, et al. , “Single-\nshot all-optical switching of magnetization in Tb/Co multilayer-based elec-\ntrodes,” Scientific reports 10, 1–8 (2020).\n23M. Beens, M. L. Lalieu, A. J. Deenen, R. A. Duine, and B. Koop-\nmans, “Comparing all-optical switching in synthetic-ferrimagnetic multi-\nlayers and alloys,” Physical Review B 100, 220409 (2019).\n24Y . Xu, M. Deb, G. Malinowski, M. Hehn, W. Zhao, and S. Mangin, “Ul-\ntrafast magnetization manipulation using single femtosecond light and hot-\nelectron pulses,” Advanced Materials 29, 1703474 (2017).\n25T. H. Pham, J. V ogel, J. Sampaio, M. Va ˇnatka, J.-C. Rojas-Sánchez,\nM. Bonfim, D. Chaves, F. Choueikani, P. Ohresser, E. Otero, et al. , “Very\nlarge domain wall velocities in Pt/Co/GdOx and Pt/Co/Gd trilayers with\nDzyaloshinskii-Moriya interaction,” EPL (Europhysics Letters) 113, 67001\n(2016).\n26P. Li, T. J. Kools, B. Koopmans, and R. Lavrijsen, “Ultrafast racetrack\nbased on compensated Co/Gd-based synthetic ferrimagnet with All-Optical\nSwitching,” Advanced Electronic Materials 9, 2200613 (2023).\n27L. Wang, H. Cheng, P. Li, Y . L. van Hees, Y . Liu, K. Cao, R. Lavri-\njsen, X. Lin, B. Koopmans, and W. Zhao, “Picosecond optospintronic\ntunnel junctions,” Proceedings of the National Academy of Sciences 119,\ne2204732119 (2022).\n28H. Windischmann, “Intrinsic stress in sputter-deposited thin films,” Critical\nReviews in Solid State and Material Sciences 17, 547–596 (1992).\n29K. Twarowski and H. Lachowicz, “Magnetostriction and anisotropy of\namorphous Gd-Co RF sputtered thin films,” Journal of Applied Physics 50,\n7722–7724 (1979).\n30Z. Chen, X. Shi, X. Liu, X. Chen, Z. Zhang, and W. Mi, “Modulating\nsaturation magnetization and topological Hall resistivity of flexible ferri-\nmagnetic Mn4N films by bending strains,” Journal of Applied Physics 132,\n233906 (2022).\n31M. Zheng, P. Guan, and H. Fan, “Mechanically enhanced magnetism\nin flexible semitransparent CuFe2O4/mica epitaxial heterostructures,” Ap-\nplied Surface Science 584, 152586 (2022).32A. Tavassolizadeh, K. Rott, T. Meier, E. Quandt, H. Hölscher, G. Reiss,\nand D. Meyners, “Tunnel magnetoresistance sensors with magnetostrictive\nelectrodes: Strain sensors,” Sensors 16, 1902 (2016).\n33A. M. Sahadevan, R. K. Tiwari, G. Kalon, C. S. Bhatia, M. Saeys, and\nH. Yang, “Biaxial strain effect of spin dependent tunneling in MgO mag-\nnetic tunnel junctions,” Applied Physics Letters 101, 042407 (2012).\n34Q. Wang, J. Domann, G. Yu, A. Barra, K. L. Wang, and G. P. Carman,\n“Strain-mediated spin-orbit-torque switching for magnetic memory,” Phys-\nical Review Applied 10, 034052 (2018).\n35G. Masciocchi, M. Fattouhi, A. Kehlberger, L. Lopez-Diaz, M.-A. Syskaki,\nand M. Kläui, “Strain-controlled domain wall injection into nanowires for\nsensor applications,” Journal of Applied Physics 130, 183903 (2021).\n36C. Chappert, A. Fert, and F. N. Van Dau, “The emergence of spin electron-\nics in data storage,” Nature materials 6, 813–823 (2007).\n37B. Tudu and A. Tiwari, “Recent developments in perpendicular magnetic\nanisotropy thin films for data storage applications,” Vacuum 146, 329–341\n(2017).\n38M. Johnson, P. Bloemen, F. Den Broeder, and J. De Vries, “Magnetic\nanisotropy in metallic multilayers,” Reports on Progress in Physics 59, 1409\n(1996).\n39F. Den Broeder, W. Hoving, and P. Bloemen, “Magnetic anisotropy of mul-\ntilayers,” Journal of magnetism and magnetic materials 93, 562–570 (1991).\n40K. Kyuno, J.-G. Ha, R. Yamamoto, and S. Asano, “Theoretical study on\nthe strain dependence of the magnetic anisotropy of X/Co (X= Pt, Cu, Ag,\nand Au) metallic multilayers,” Journal of applied physics 79, 7084–7089\n(1996).\n41S. Hashimoto, Y . Ochiai, and K. Aso, “Perpendicular magnetic anisotropy\nand magnetostriction of sputtered Co/Pd and Co/Pt multilayered films,”\nJournal of applied physics 66, 4909–4916 (1989).\n42K. Twarowski, H. Lachowicz, M. Gutowski, and H. Szymczak, “On the\norigin of the perpendicular anisotropy and magnetostriction in amorphous\nRF sputtered Gd Co films,” physica status solidi (a) 63, 103–108 (1981).\n43G. Masciocchi, J. W. van der Jagt, M.-A. Syskaki, A. Lamperti, N. Wolff,\nA. Lotnyk, J. Langer, L. Kienle, G. Jakob, B. Borie, et al. , “Control of\nmagnetoelastic coupling in Ni/Fe multilayers using He+ ion irradiation,”\nApplied Physics Letters 121, 182401 (2022).\n44T. J. Kools, M. C. van Gurp, B. Koopmans, and R. Lavrijsen, “Magne-\ntostatics of room temperature compensated Co/Gd/Co/Gd-based synthetic\nferrimagnets,” Applied Physics Letters 121, 242405 (2022).\n45G. Bertero, T. Hufnagel, B. Clemens, and R. Sinclair, “TEM analysis of\nCo-Gd and Co-Gd multilayer structures,” Journal of materials research 8,\n771–774 (1993).\n46B. Zhang, K. M. Krishnan, C. Lee, and R. Farrow, “Magnetic anisotropy\nand lattice strain in Co/Pt multilayers,” Journal of applied physics 73, 6198–\n6200 (1993).\n47D. B. Gopman, C. L. Dennis, P. Chen, Y . L. Iunin, P. Finkel, M. Staruch,\nand R. D. Shull, “Strain-assisted magnetization reversal in Co/Ni multi-\nlayers with perpendicular magnetic anisotropy,” Scientific reports 6, 1–8\n(2016).Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nSupplementary material - Strain effects on magnetic compensation and spin\nreorientation transition of Co/Gd synthetic ferrimagnets\n(Dated: 3 April 2023)\n1arXiv:2303.15191v4 [cond-mat.mtrl-sci] 31 Mar 2023Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nS1 - Magnetostatics model for the Spin Reorientation Transition\nIn the expression of the free energy, a term describing the effect of intermixing at the multilayer\nsystem interfaces is added. The expression is\nEmix=1\n2µ0/integraltexta0x\n0M2\nCo+ (MGdexp(−q/λGd))2dq=\n1\n2µ0a0M2\nCox+1\n4µ0λGdM2\nGd/parenleftBig\n1−exp/parenleftBig\n−2a0x\nλGd/parenrightBig/parenrightBig\n.(S.1)\nTherefore, the total energy Etot=−EK−Emix−KME+Ed,Co+Ed,Gd, including all the terms\nwill be:\nEtot=−Ks+∆K/parenleftBig\n1−exp/parenleftBig\n−2x\nλK/parenrightBig/parenrightBig\n−KME−1\n2µ0a0M2\nCox\n−1\n4µ0λGdM2\nGd/parenleftBig\n1−exp/parenleftBig\n−2a0x\nλGd/parenrightBig/parenrightBig\n+1\n2µ0M2\nCoy\n+1\n4µ0M2\nGdλGd/parenleftBig\n1−exp/parenleftBig\n−2x\nλGd/parenrightBig/parenrightBig\n.(S.2)\nHere xandyare, respectively, the Gd and Co thicknesses in the phase diagram of Fig. S1. The\nvalue of the parameters used in our model are listed in Table S I .\nParameter Value Description\nKs 1.7 mJ/m2Interfacial anisotropy (from exp.)\nKME 0.02 mJ/m2Magnetoelastic anisotropy (from exp.)\nMCo 1.3 MA/m Cobalt magnetization (from exp.)\nMGd 1.4 MA/m Gadolinium magnetization at Co/Gd interface (from Ref.1)\na0 0.13 (-) Growth parameter of intermixing region (from exp.)\nλK 0.51(15) nm Change of PMA energy characteristic decay length (Fit parameter)\nλGd 0.59(22) nm Gd magnetization decay characteristic decay length (Fit parameter)\n∆K 3.96(41) ×10−4J/m2Change of PMA energy (Fit parameter)\nTABLE S I: Parameters used in the model for the magnetostatics of uncompensated Co/Gd synthetic ferrimagnets\nused for the calculations of the SRT. The term KMEis considered zero for when external strain is not applied to the\nsample.\nThe values of λK,λGdand∆Kare instead determined using a numerical fit and are reported\nin Table S I. To fit this equation to the phase diagram obtained experimentally, it is convenient to\n2Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nfind the Co-thickness (y) where the anisotropy energy ( Etot) is equal to zero (spin reorientation\ntransition, SRT). Solving Eq. S.2 for y gives:\ny0(x) =2\nM2\nCoµ0/parenleftBig\n−/parenleftBig\nKs−∆K/parenleftBig\n1−exp/parenleftBig\n−x\nλK/parenrightBig/parenrightBig/parenrightBig/parenrightBig\n−1\n2µ0a0M2\nCox\n−1\n4µ0λGdM2\nGd/parenleftBig\n1−exp/parenleftBig\n−2a0x\nλGd/parenrightBig/parenrightBig\n+1\n4µ0M2\nGdλGd/parenleftBig\n1−exp/parenleftBig\n−2x\nλGd/parenrightBig/parenrightBig\n.(S.3)\nNote that for determining the fit parameters, the measurements were taken without externally\napplied strain. Accordingly, the magnetoelastic energy term KMEis set to zero in Eq. S.3. The\nexperimental data used for the numerical fit are reported in Fig. S1. The sample used for the\nnumerical fit, explores a wide thickness range (Gd from 0 to 6 nm) in a double wedge fashion\nto improve accuracy. Consequently, the dimensions of this sample exceed the 1x1 cm size of\nthe bending device. For this reason, single wedge samples have been deposited for the strain-\ndependent study.\nFIG. S 1: Values for the SRT obtained experimentally on a Ta(4nm)/Pt(4)/Co(tCo)/Gd(tGd)/TaN (4)sample and\nused to extract the fitting parameters in Eq. S.3.\nThe feature around tGd=2 nm in Fig. S1 is not captured by out toy model, and might be due to\nthe additional intermixing caused during sputtering, not included in Eq. S.2.\n3Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nS2 -Application of strain\nTo obtain information about the magnetoelastic properties of the material, the substrate was\nbent mechanically with a 3 point bending sample holder, as shown schematically in Fig. S2 (a).\nA square sample of 1 by 1 cm is vertically constrained on two sides and pushed uniformly from\nbelow by a cylinder that has an off-centered rotation axis. The device generates a tensile strain\nin the plane of the sample up to 0 .1 % when the cylinder is rotated by 90◦. The strain is mostly\nuniaxial and has been measured with a strain gauge on the substrate surface.\nFIG. S 2: (a) schematic of the three point bending method used to externally strain the sample. The strain is mostly\nuniaxial along the xdirection. (b) hysteresis loops measured before (blue) and during (red) application of in-plane\nstrain for a sample of Pt/Co(1.85 nm)/Ta. The area highlighted in red corresponds to the magnetoelastic energy in the\nstrained system. The magnetic field was applied along the OOP direction ( z).\nMagnetic hysteresis loops are recorded before and after the application of the tensile strain\nand are used to estimate the magnetoelastic anisotropy. As previously reported2,3the magnetic\nanisotropy Ke f fis linked to the energy stored in the magnetization curves. For example the PMA\nenergy is given by the area enclosed between the magnetic loops measured with field along IP and\nOOP direction. If then the strain in the film is non-zero, the magneto-elastic coupling contributes\nin principle to the effective anisotropy. Two hysteresis loops measurements, before and after the\napplication of strain, are sufficient to estimate KME. Indeed the total anisotropy of the system\nisKe f f=KsandKe f f=Ks+KMEbefore and after the application of strain, respectively. The\nmagnetoelastic anisotropy KME=−3\n2λsYεis linked to reversible part of the hysteresis loops (close\nto the saturation) according to\n4Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nKME=Ms∆E=−3\n2λsYε (S.4)\nwhere ∆Eis the anisotropy energy measured by the difference in area below the strained and\nunstrained curves, εis the strain λsis the magnetostriction and Y is the Young’s mudulus of the\nmaterial. In our case ε=0.1% and Y=200 GPa. ∆Ecorresponds to the reversible part, i.e. the red\nmarked area in Fig. S2 (b). The value of magnetoelastic anisotropy was calculated using the value\nof saturation magnetization ( Ms) of the stack taken from literature and reported in Table S I.\n5Suppl. material - Strain effects on magnetic compensation and spin reorientation transition of ...\nREFERENCES\n1Kools, T. J., van Gurp, M. C., Koopmans, B., and Lavrijsen, R. (2022). Magnetostatics of room\ntemperature compensated Co/Gd/Co/Gd-based synthetic ferrimagnets. Applied Physics Letters,\n121(24), 242405.\n2Johnson, M. T., Bloemen, P. J. H., Den Broeder, F. J. A., and De Vries, J. J. (1996). Magnetic\nanisotropy in metallic multilayers. Reports on Progress in Physics, 59(11), 1409.\n3Baril, L., Gurney, B., Wilhoit, D., Speriosu, V . (1999). Magnetostriction in spin valves. Journal\nof Applied Physics, 85(8), 5139-5141.\n6" }, { "title": "1607.02358v3.Control_of_magnon_photon_coupling_strength_in_a_planar_resonator_YIG_thin_film_configuration.pdf", "content": "arXiv:1607.02358v3 [cond-mat.mtrl-sci] 24 Nov 2016Control of magnon-photon coupling strength in a planar reso nator/YIG thin film\nconfiguration\nV. Castel,1A. Manchec,2and J. Ben Youssef3\n1)T´ el´ ecom Bretagne, Technopole Iroise-Brest, CS83818, 29 200 Brest,\nFrance.\n2)Elliptika (GTID), 29200 Brest, France.\n3)Universit´ e de Bretagne occidentale, Laboratoire de Magn´ etisme de Bretagne CNRS,\n29200 Brest, France\n(Dated: 26 July 2021)\nA systematic study of the coupling at room temperature between f erromagnetic res-\nonance (FMR) and a planar resonator is presented. The chosen ma gnetic material is\na ferrimagnetic insulator (Yttrium Iron Garnet: YIG) which is positio ned on top of a\nstop band (notch) filter based on a stub line capacitively coupled to a 50 Ω microstrip\nline resonating at 4.731 GHz. Control of the magnon-photon couplin g strength is dis-\ncussed interms ofthe microwave excitation configurationandtheY IG thickness from\n0.2 to 41 µm. From the latter dependence, we extract a single spin-photon co upling\nof g0/2π=162±6 mHz and a maximum of an effective coupling of 290 MHz.\nKeywords: Cavity spintronic, quantum detector, Yttrium Iron Ga rnet, notch filter,\nmagnon-photon coupling\n1I. INTRODUCTION\nA recent field, known as cavity spintronics1,2, is emerging from the progress of spintronics\ncombined with the advancement in Cavity Quantum Electrodynamics ( QED) and Cavity\nPolaritons3,4. Cavity QED allows the use of coherent quantum effects for quantu m informa-\ntion processing and offers original possibilities for studying the stro ng interaction between\nlight and matter in a variety of solid-state systems5–7. A superconducting two-level system\nis quantum coherently coupled to a single microwave photon and an an alogy to spintronics\n(spin two-level system) has been made. The high spin density of the ferromagnet used in\nRef.8,9hasmade it possible to create a strongly coupled magnonmode. Magn on-photoncou-\npling has been investigated in several experiments at room tempera ture where a microwave\nresonator (three-dimensional cavity10–18and planar configuration19–21) was loaded with a\nferrimagnetic insulator such as the Yttrium Iron Garnet (YIG, thin film and bulk). A study\non a transition metal like Py (structured thin film) coupled with a Split R ing Resonator\n(SRR) was done by Gregory et al.22in order to demonstrate the possibility to achieve YIG-\ntype functionalities and to overtake the working frequency limitatio n of YIG. More recently,\nL. Bai et al.15have developed an electrical method to detect magnons coupled wit h photons.\nThis method has been established by placing a hybrid YIG/Pt system in a microwave cavity\nshowing distinct features not seen in any previous spin pumping expe riments but already\npredicted by Cao et al.23.\nII. COMPACT DESIGN DESCRIPTION\nThe main objective of the present paper is to demonstrate the con trol of a magnon-\nphoton coupling regime at room temperature in a compact design bas ed on a stub line\ncoupled with a microstrip with YIG thin film. Control of a magnon-phot on coupling regime\nin such configuration offers manifold opportunities in the developmen t of integrated spin-\nbased microwave applications, such as a sensitive reconfigurable st op-band filtering function.\nInsteadofusingaSplitRingResonator(SRR)configuration,thech oicewasmadetostudy\ntheYIGthickness dependence ofthecoupling regime withastub lineg eometry forwhich the\nmicrowaveexcitationofthemagneticmediumissimplified. Figure1(a)r epresentsthesketch\nof the experimental setup based on the stub/YIG film system excit ed by a microwave signal\n2/s76\n/s115/s119\n/s115/s49/s103\n/s119\n/s115/s50\n/s119/s40/s99/s41/s40/s98/s41\n/s83/s32/s112/s97/s114/s97/s109/s101/s116/s101/s114/s115/s32/s91/s100/s66/s93/s32/s83\n/s49/s49/s32/s40/s77/s41\n/s32/s83\n/s50/s49/s32/s40/s77/s41\n/s32/s83\n/s49/s49/s32/s40/s83/s41\n/s32/s83\n/s50/s49/s32/s40/s83/s41\n/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32/s91/s71/s72/s122/s93/s40/s97/s41\n/s89/s73/s71/s47/s71/s71/s71\n/s72/s48\n/s53\n/s49/s53/s49/s48\nFIG. 1. (a) Experimental setup: A Vector Network Analyzer (V NA) is connected to a 50 Ω\nmicrostrip line which is capacitively coupled to the resona tor. Numbers from 0 to 15 are referred\nto the YIG sample position (an example is given for which the c enter of the sample is placed\nat x=10 mm=0.25 λ). (b) Dimension of the microwave stop band resonator configu ration. (c)\nFrequency dependence of S parameters measured (M) and simul ated (S) by CST simulation of the\nempty resonator (without YIG sample).\nunder anin-plane static magnetic field, H, at θ=0◦.θis defined by the angle formed between\nH and the stub line. P1 and P2 correspond to the 2 ports of the VNA f or which a TSOM\ncalibration were realized (including cables). The frequency range is fi xed from 3 to 6 GHz at\nan microwave power of P=-10 dBm. The circuit is fabricated on a pre- metallized (double-\nsided 25 µm copper coating) ROGERS substrate (3003) presenting a relative permittivity\nofεr=3 and losses tan δ=2×10−3(dimension are shown in Fig. 1 (b)). The narrow stop-\nband (notch) resonator configuration is based on a main 50 Ω micros trip line (W=1.23 mm)\ncoupled by a gap of g=150 µm to an open circuited half wavelength stub (L s=15.52 mm,\nWs1=280µm and W s2=1.0 mm). It has been designed with an attenuation of 13 dB at 4.75\nGHz and 80 MHz of bandwidth. The frequency dependence of the S1 1 and S21 parameters\n(measured andsimulated) oftheempty resonatorareshown inFig. 1(c). TheS21resonance\npeak has a half width at half maximum (HWHM) ∆F HWHMof 32 MHz indicating that the\ndampingoftheresonator(workingatthefrequency F0)isβ=∆FHWHM/F0=1/2Q=6.8 ×10−3.\nThis leads to a quality factor Q of 74. The latter definitions of the Q fa ctor and β(used\n3in the following discussion) are based on the expression extracted f rom recent studies in the\nfield of magnon-photon coupling (2D20,21and 3D cavities13–18). Nevertheless, this definition\ndoes not reflect properly the electrical performances of our not ch filter which are defined by\nQ0=F0//bracketleftbig\n∆FS21\n−3dB(1−S11F0)/bracketrightbig\n=153.\nSingle-crystal Y 3Fe5O12(YIG) samples from 0.2 to 41 µm were elaborated by Liquid\nPhase Epitaxy (LPE) on top of a 500 µm thick GGG substrate in the (111) orientation.\nYIG samples have been cut in rectangular shape (4 mm ×7 mm) and placed on the stub line\nas shown in Fig. 1 (a) with the crystallographic axis [1,1, ¯2] parallel to the planar microwave\nfield generated by the stub line. The magnetic losses (Gilbert damping parameter, α) of the\nsetofYIGsamplewereinvestigatedbyFMRmeasurementsusingahig hlysensitivewideband\nresonance spectrometer within a range of 4 to 20 GHz. Measureme nts were carried out at\nroom temperature with a static magnetic field applied in the plane of YI G samples. These\ncharacterizations have given rise to a parameter α≤2×10−4for the set of samples which is\nin agreement with previous studies24.\nIII. RESULTS AND DISCUSSION\nWe first studied the magnitude of the coupling strength as a functio n of the position\nof a 9µm YIG sample on top of the planar resonator, as shown in Fig. 1 (a). F igures\n2 (a) to (c) illustrate the dependence of the resonator features (at H=0 Oe), such as the\nresonant frequency F0, linewidth ∆F HWHM, damping of the resonator β(Q factor), and\nthe dependence of S11 and S21 at the resonant frequency F0.F0can be tuned from 4.35\nto 4.715 GHz (tuning of 8.4 %) and presents a maximum at x=0.25 λwhich is closed to\nthe resonant frequency of the empty resonator (represented by horizontal dash dot lines).\nNote that the YIG position corresponds to the center of the samp le as illustrated in Fig.\n1 (a). The wavelength is defined by λ=λ0√εeff, whereεeffcorresponds to the effective\npermittivity. The configuration of the notch filter (open circuit (OC ) at x=0.5 λ) induced\nnecessarily the definition of short circuit (SC) at x=0.25 λwhich explains the limited impact\nof YIG (at this position) on the resonator features. Introductio n of a YIG layer ( εr=15\nand losses tan δ=2×10−4) on CST simulation make it possible to correctly reproduce this\ndependence at zero field (solid black line) which is attributed to the mo dification of the\neffective permittivity. Here, only the electrical properties of the Y IG sample were taken into\n4FIG. 2. (a) to (c): YIG sample position dependence (at H=0 Oe) of (a)F0and ∆F HWHM,\n(b) Q factor and losses, (c) S11 and S21 at the resonant freque ncyF0. Horizontal dash dot\nlines correspond to parameters extracted from the empty res onator whereas the vertical dash dot\nline illustrates the position of YIG (respect to the center) at x=0.25 λas shown in Fig. 1 (a).\n(d) and (e): Signature of the coupling. Frequency dependenc e of the magnitude in dB of S21\n(d) and the associated phase φS21(e) for a YIG position at x=0.25 λ. Solid blue and red line\ncorrespond respectively to the response at H=0 Oe and at H=H RES.F1andF2(represented by\nvertical red dash lines) correspond to hybridized mode freq uencies whereas g eff/2πcorresponds\nto the coupling strength parameter. (f) Dependence of g eff/2π(measured, black triangles) and\nmicrowave field amplitude (CST simulation, solid red line) a s function of the YIG sample position.\nAll measurements have been carried on at room temperature on a YIG sample which presents\na thickness of (9 µm). The inset shows the spatial distribution of the microwav e magnetic field\nsimulated at F0.\naccount. Contrary to a ferromagnetic conductor, such as an ex tended thin film of Permalloy\n(Py, NiFe)22, no eddy current shielding effect of YIG on the stub line was observe d. YIG is\na ferrimagnetic insulator with a band gap of 2.85 eV and the high quality YIG samples used\n5in this study allow the reduction of negative impacts on the planar res onator. As shown\nin Fig. 2 (a), ∆F HWHMreaches a maximum of 34 MHz at x=0.3 λ, which represents an\nenhancement of only 2 MHz with respect to the empty resonator (a reduction of 5 MHz\nbeing achievable at x=0.45 λ). Slightly changes in the Q factor ( β) from 70 to 77 (6.4 to\n7.2×10−3) were obtained. In the meantime, attenuation represented by th e S11 parameter\natF0are closed to the value extracted from the empty resonator from x=0.4 to 0.3 λ(-2.35\ndB).\nFor each position of the YIG sample, measurement at room tempera ture of the frequency\ndependence of S parameters (magnitude and phase) at P=-10 dBm was done with respect\nto the applied magnetic field. Figure 2 (d) and (e) represent, respe ctively, the frequency\ndependence of S21 and Φ S21of the notch/YIG system at x=0.25 λ(as shown in Fig. 1\n(a)). Solid (dash) blue and red lines are associated with the experime ntal (analytic solution\nfrom Eq. (3) from Ref.18) response under an applied magnetic field of H=0 Oe and H=H RES,\nrespectively. TheFMRandthenotchfilterinteractbymutualmicro wavefields, generatedby\nthe oscillating currents in the stub and the FMR magnetization prece ssion which led to the\nfollowing features observed in Fig. 2 (d) and (e): (i) Hybridization of resonances (magnitude\nand phase25), (ii) Annihilation of the resonance at F0, and (iii) Generation of two resonances\natF1andF2. At the resonant condition H=H RES, the frequency gap, Fgap, between F1and\nF2is directly linked to the coupling strength of the system ( Fgap/2=geff/2π). Several models\ncan be used to analyze the hybridized mode frequency F1andF2in the system. Recently,\nHarder et al.18have examined the accuracy to describe the microwave transmissio n line\nshape of a cavity/YIG system through three different models: cou pled harmonic oscillators,\ndynamic phase correlation, and microscopy theory. Here, the ana lysis has been focussed on\nthe harmonic coupling model for which we can define theupper ( F2) andlower ( F1) branches\nby:\nF1,2=1\n2/bracketleftbigg\n(F0+Fr)±/radicalBig\n(F0−Fr)2+k4F2\n0/bracketrightbigg\n(1)\nTheFMRfrequency, Fr, ismodelledbytheKittelequation, Fr=γ\n2πµ0/radicalbig\nH(H+Ms), which\ndescribes the precession frequency of the uniform mode (without taking into account spin\nwave distribution) in an in-plane magnetized ferromagnetic film. The p arameter kused in\nEq. 1 corresponds to the coupling strength which is linked to the exp erimental data g eff/2π\n6by the following equation18:Fgap=F2−F1=k2F0. As shown in Fig. 2 (f), sensitive\ncontrol of g eff/2π(and thus k) can be achieved by adjusting the YIG sample position on\nthe resonator from 54 MHz ( k=0.1573) at x=0.45 λto 127 MHz ( k=0.2315) at x=0.25 λ.\nIn order to understand the dependence of g eff/2πon the YIG position, CST simulations\nwere carried out in order to determine the microwave field ( hMW) generated at each position\n(represented in solid red line). It ends up that hMWfollows exactly the same trend of the\ncoupling factor, in agreement with the fact that the effective coup ling strength depends on\nthe mutual microwave field interaction between the FMR and the stu b line. The latter\ndependence is defined by the following equation9,12:\ngeff\n2π=η\n4πγe/radicalbigg\n/planckover2pi1ω0µ0\nVc√\nN, (2)\nwhereγeistheelectrongyromagneticratioof2 π×28.04GHz/T, µ0isthepermeability ofthe\nvacuum, Vccorresponds to the volume of the cavity, and Nis the total number of spins. The\ncoefficient η≤1 describes the spatial overlap and polarization matching conditions between\nthe microwave field and the magnon mode. In agreement with Zhang e t al.12(Appendix\nA), we demonstrated the dependence of g eff/2πas function of the spatial distribution of the\nmicrowave magnetic field along the stub line which is maximum at x=0.25 λ(short circuit).\nTABLE I. Notch/YIG configuration versus SRR & cavity/YIG sys tems\nRef. β[10−3]geff/2π[MHz]F0[GHz] k\n151.8 80 10.506 0.1234\n160.708 31.8 9.650.0812\n172.3 65 10.847 0.1095\n180.3 31.5 10.556 0.0773\n13,141.92 130 3.5350.2712\n209.85 270 3.20.4108\n215.04 63 4.960.1594\nThis work[1]6.89 127 4.7160.2315\nThis work[2]6.89 290 4.7190.3508\nTable I gives a picture of recent work on the determination and cont rol of magnon-\nphoton coupling regimes in SRR20,21and cavity13–18/YIG systems. Ref.15–17correspond to\n7the research field associated with the electrical detection of magn ons coupled with photons\nvia combined phenomena in a hybrid YIG/Pt system. Despite the fact that these later\nstudies have been realized in a cavity, insertion of a hybrid stack inclu ding a highly electrical\nconductor induced an enhancement of the intrinsic loss rate β(factor of 515to 1216). The\nvalueofkobtainedattheoptimizedpositionatx=0.25 λissignificantlyhigherthanRef.15–18\nand comparable to Ref.13,14,21but still much smaller than the value obtained by Bhoi et al.20.\nIt should be noted that the normalization of kby the intrinsic loss rate βchanges the latter\ncomparison drastically.\nNext, the dependence of the coupling strength between the FMR a nd the notch filter\nwas investigated with respect to the YIG thickness from 0.2 to 41 µm. YIG samples were\nplaced at the optimized position which has been determined previously (x=0.25 λ). This\nparticular position gives an access to the highest coupling (determin ed at P=-10 dBm) and\npresents the best compromise in terms of the electrical performa nce of the notch filter.\nSample position was adjusted by tracking F0at H=0 Oe ( F0=4.715±0.002 GHz). It should\nbe noted that no dependence of the insertion rate of the resonat or (β=6.89±0.01 10−3) and\nattenuation (S11 F0=-2.45±0.04 dB) have been observed with respect to the YIG thickness.\nAs shown in Fig. 3 (a), we demonstrated a strong coupling regime via t he anti-crossing\nfingerprint. A good agreement of F1,2based on Eq. 1 (solid lines) with experimental data\nis obtained for the various YIG thicknesses. The color plot in Fig. 3 (a ) is associated with\nthe S21 parameter for which the dark area corresponds to a magn itude of -10 dB. This\nrepresentation underlines the complexity of the response by incre asing the YIG thickness\nfrom 0.2 to 41 m, well illustrated by the additional anti-crossing signa ture between 0.75\nand 0.90 kOe (upper resonance). In the following discussion, the ex traction of the coupling\nfactor is only based on the uniform mode without taking into account the dispersion relation\nof spin waves. Figure 3 (b) represents the frequency dependenc e of the transmission spectra\nfor the notch/YIG system at the resonant condition for which the effective coupling was\nextracted. Control of the frequency gap can be achieved from 5 9 to 581 MHz through an\nenhancement of the YIG thickness from 0.2 to 41 µm, respectively. Parameters associated\nwith the thicker YIG are summarized in Tab I (last row).\nThe originality of this study is described in Fig. 3 (c) which represents the dependence of\nthe effective coupling g eff/2πas a function of the square root of the YIG volume interacted\nwith the 1 mm width microwave resonator (V=4 mm ×1 mm×YIGthickµm). Cao et al.23\n8FIG. 3. Control of the coupling strength as function of the YI G thickness. (a) Magnetic field\ndependence of the frequency: Observation of the strong coup ling regime via the anti-crossing\nfingerprint. The color map is associated to the response of th e thicker YIG sample (41 µm). (b)\nFrequency dependence of S21 at the resonant condition for va rious YIG thickness (0.2, 9, and 41\nµm). (c) Coupling strength of the Kittel mode to the microwave resonator mode as a function of\nthe square root of the YIG volume. Colored triangles corresp ond to the dispersion of g eff/2πfrom\nFig. 2 (f). The inset represents the YIG thickness dependenc e ofk. All measurements were done\nat room temperature and at x=8 mm as shown in Fig. 1 (a).\nshows that the filling factor of magnetic medium in a cavity can be used as a measure of\nthe total number of spins, N. The large effective coupling strength is due to the large\nspin density of YIG, ρs=2.1×1022µBcm−3(µB; Bohr magneton). The linear fit of the\ndependence presented in Fig. 3 (c) gives rise to a slope of 742 ±29 MHz mm3/2which\n9makes it possible to extract the single spin-photon coupling g 0/2π=162±6mHz based on the\nfollowing equation9,16geff=g0√\nN. Tabuchi et al.9demonstrate a good agreement between\nthe evaluation of the single-spin coupling strength from the fitting ( 39 mHz) and theory\nderived from the quantum optics community (38 mHz). The latter va lue is calculated from\nEq.2 by taking the coefficient η= 1. The higher value of g 0/2πobtain in the present work\nis mainly due to the compactness of our resonator. A rough estimat ion of the volume of our\nnotch filter working at F0=4.750 GHz can be done by using a one dimensional transmission-\nline cavity26which defines the volume as Vc=πr2λ/2. By assuming r=0.5 mm (distance\nbetween the feed line and the ground), λ=c/(√εeffF0), andη= 1, we evaluate g 0/2π=177\nmHzwhenthecavityiscompletely filledbyair( εeff=εr=1)which isclosed totheextracted\nvalue of g 0/2πfrom the fitting. Nevertheless, this value does not reflect the fac t that the\ncavity is nonuniformly filled with other dielectric materials such as the s ubstrate, GGG, and\nYIG. An enhancement of g 0/2πfrom 177 to 219 mHz can be achieved by taking into account\nan effective permittivity of εeff=2.449. The latter value is determined27for the notch filter\nloaded with a YIG film at x=0.25 λwhich induced a diminution of F0from 4.750 to 4.715\nGHz.\nWe have demonstrated the presence of a strong coupling regime via the anti-crossing\nfingerprint of the FMR from an magnetic insulator and a planar reson ator. Control of\nthe coupling with respect to the YIG thickness from 0.2 to 41 µm makes it possible the\ndetermination of the single spin photon coupling of our system at roo m temperature. We\nhave found that g 0/2πin a thin film configuration is equal to 162 ±6 mHz. In the meantime,\nwe demonstrate an effective coupling strength of 290 MHz for the t hicker YIG. Improvement\non insertion losses of the planar resonator can be achieved in order to be more competitive\nregarding the 3D cavity system by changing the design of the reson ator (SRR, array of\nSRR, enhancement of the capacitive coupling) and/or by using a low lo ss substrate ( <\n10−2). Other tuning channel of the coupling strength such as microwav e power or spin wave\ndispersion might help for the realization of YIG-based devices.\nREFERENCES\n1O. O. Soykal and M. E. Flatt´ e, “Strong field interactions between a nanomagnet and a\nphotonic cavity,” Phys. Rev. Lett. 104, 077202 (2010).\n102C. M. Hu, “Dawn of cavity spintronics,” Physics in Canada 72, 76 (2016).\n3R. W. Sanders, V. Jaccarino, and S. Rezende, “Magnetic polariton , impurity mode en-\nhancement, and superradiance effects infef2,” SolidStateCommun ications28, 907(1978).\n4D. L. Mills and E. Burstein, “Polaritons: the electromagnetic modes o f media,” Reports\non Progress in Physics 37, 817 (1974).\n5C. Cohen-Tannoudji, “Atoms in electromagnetic fields,” World Scient ific Series onAtomic,\nMolecular and Optical Physics 3(2004).\n6R. Laflamme, E. Knill, D. G. Cory, E. M. Fortunato, T. Havel, C. Mique l, R. Martinez,\nC. Negrevergne, G. Ortiz, M. a. Pravia, Y. Sharf, S. Sinha, R. Som ma, and L. Viola,\n“Introduction to nmr quantum information processing,” arXiv (200 2).\n7A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, J. Majer, S. Kumar , S. M. Girvin, and\nR. J. Schoelkopf, “Strong coupling of a single photon to a supercon ducting qubit using\ncircuit quantum electrodynamics,” Nature 431, 1 (2004).\n8H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Ma rx, R. Gross, and\nS. T. B. Goennenwein, “High cooperativity in coupled microwave reso nator ferrimagnetic\ninsulator hybrids,” Phys. Rev. Lett. 111, 127003 (2013).\n9Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Na kamura, “Hy-\nbridizing ferromagnetic magnons and microwave photons in the quan tum limit,” Phys.\nRev. Lett. 113, 083603 (2014).\n10L. Kang, Q. Zhao, H. Zhao, and J. Zhou, “Magnetically tunable nega tive permeability\nmetamaterial composed by split ring resonators and ferrite rods,” Opt. Express 16, 8825\n(2008).\n11J. N. Gollub, J. Y. Chin, T. J. Cui, and D. R. Smith, “Hybrid resonant p henomena in a\nSRR/YIG metamaterial structure.” Opt. Express 17, 2122 (2009).\n12X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, “Strongly coupled ma gnons and cavity\nmicrowave photons,” Phys. Rev. Lett. 113, 156401 (2014).\n13N. J. Lambert, J. a. Haigh, and a. J. Ferguson, “Identification of spin wave modes in\nyttrium iron garnet strongly coupled to a co-axial cavity,” Journal of Applied Physics\n117, 053910 (2015).\n14J. A. Haigh, N. J. Lambert, A. C. Doherty, and A. J. Ferguson, “D ispersive readout\nof ferromagnetic resonance for strongly coupled magnons and mic rowave photons,” Phys.\nRev. B91, 104410 (2015).\n1115L. Bai, M. Harder, Y. P. Chen, X. Fan, J. Q. Xiao, and C.-M. Hu, “Spin pumping\nin electrodynamically coupled magnon-photon systems,” Phys. Rev. Lett.114, 227201\n(2015).\n16H.Maier-Flaig, M. Harder, R.Gross, H. Huebl, andS.T. B.Goennen wein, “Spinpumping\nin strongly coupled magnon-photon systems,” Phys. Rev. B 94, 054433 (2016).\n17L. Bai, K. Blanchette, M. Harder, Y. P. Chen, X. Fan, J. Q. Xiao, an d C. M. Hu, “Control\nof the magnon-photon coupling,” IEEE Transactions on Magnetics 52, 1 (2016).\n18M. Harder, L. Bai, C. Match, J. Sirker, and C. Hu, “Study of the ca vity-magnon-polariton\ntransmission line shape,” Science China Physics, Mechanics & Astrono my59, 117511\n(2016).\n19G. B. G. Stenning, G. J. Bowden, L. C. Maple, S. A. Gregory, A. Spo sito, R. W. Eason,\nN. I. Zheludev, and P. A. J. de Groot, “Magnetic control of a meta -molecule,” Opt.\nExpress21, 1456 (2013).\n20B. Bhoi, T. Cliff, I. S. Maksymov, M. Kostylev, R. Aiyar, N. Venkatar amani, S. Prasad,\nand R. L. Stamps, “Study of photon-magnon coupling in a yig-film split -ring resonant\nsystem,” Journal of Applied Physics 116, 243906 (2014).\n21S. Klingler, H. Maier-Flaig, R. Gross, C.-M. Hu, H. Huebl, S. T. B. Goen nenwein,\nand M. Weiler, “Combined Brillouin light scattering and microwave absor ption study of\nmagnon-photon coupling in a split-ring resonator/YIG film system,” A pplied Physics Let-\nters109, 072402 (2016).\n22S. A. Gregory, G. B. G. Stenning, G. J. Bowden, N. I. Zheludev, an d P. A. J.\nDe Groot, “Giant magnetic modulation of a planar, hybrid metamolecu le resonance,”\nNew Journal of Physics 16, 063002 (2014).\n23Y. Cao, P. Yan, H. Huebl, S. T. B. Goennenwein, and G. E. W. Bauer, “Exchange\nmagnon-polaritons in microwave cavities,” Physical Review B 5, 094423 (2014).\n24V. Castel, N. Vlietstra, J. B. Youssef, and B. J. V. Wees, “Yttrium iron garnet thickness\nand frequency dependence of the spin-charge current convers ion in YIG / Pt systems,”\nPhys. Rev. B 90, 214434 (2014).\n25M. Harder, P. Hyde, L. Bai, C. Match, and C.-M. Hu, “Spin dynamical phase and antires-\nonance in a strongly coupled magnon-photon system,” Phys. Rev. B 94, 054403 (2016).\n26R. J. Schoelkopf and S. M. Girvin, “Wiring up quantum systems,” Natu re451, 664–669\n(2008).\n1227I. J. Bahl and D. K. Trivedi, “A designer’s guide to microstrip line,” Micr owaves16, 174\n(1977).\n13" }, { "title": "1004.4631v1.Andreev_reflection_in_ferrimagnetic_CoFe2O4_SrRuO3_spin_filters.pdf", "content": "Andreev reflection in ferrimagnetic CoFe 2O4/SrRuO 3 spin filters \n \nFranco Rigato1, Samanta Piano2,3, Michael Foerster1, Filippo Giubileo3, Anna Maria Cucolo3, and \nJosep Fontcuberta1 \n \n1Institut de Ciència de Materials de Barcel ona, CSIC, Campus UAB, Bellaterra 08193, Spain \n \n2 School of Physics and Astronomy, University of Nottingham, University Park, Nottingham, NG7 2RD, United Kingdom \n \n3Physics Department and INFM-CNR SUPERMAT Laboratory, University of Salerno, Via S. Allende, 84081 Baronissi (SA), Italy \n \n \nWe have performed point contact spectroscopy measurements on a sample constituted by a \nmetallic ferromagnetic oxide (SrRuO 3) bottom electrode and a tunnel ferrimagnetic (CoFe 2O4) \nbarrier. Andreev reflection is observed across the tunnel barrier. From the comparison of Andreev \nreflection in SrRuO 3 and across the CoFe 2O4 barrier we infer that th e ferrimagnetic barrier has a \nspin filter efficiency not larger than +13%. The observation of a moderate and positive spin filtering \nis discussed in the context of the microstructure of the barriers and symmetry-related spin filtering \neffects. I. INTRODUCTION \n \nThe spin polarization is a key quantity for the spin tronics. In the last few years, spin filters, \nconstituted by a ferro(ferri)magne tic tunnel juncti on, have emerged as promising alternative to \ncreate artificial spin polarized current sources. Ferromagnetic spin filters rely on the spin-dependent \ntransmittance of ferromagnetic tunnel barriers due to the existence of an exchange-split band gap. \nFollowing the pioneering work of Moodera et al.1 who showed spin-filtering at low temperature \nusing EuS ferromagnetic tunnel barriers, spin-filte ring has been demonstrated using perovskite \noxides2 (BiMnO 3) and more recently, spinel oxides: NiFe 2O4 (NFO)3,4 and CoFe 2O4 (CFO)5 \nbarriers. As the Curie temperature of ferrites is well above room temperature, efficient room-\ntemperature spin-polarized sources could be obta ined using these oxides. The most simple spin-\nfilter structure is formed by two metallic electrodes: a non spin-polarized current source (M) and a ferromagnetic layer (FM), acting as a spin analyzer , with a ferro(ferri)ma gnetic tunnel barrier (FI) \nbetween them. Determination of the spin-filtering efficiency (P\nFI) of the FI barrier has been \nachieved by measuring the magnetoresistance of the tunnel junction and using the Jullière model to \nderive P FI. Using this approach low-temperature values of +22% and -25% have been reported for \nNFO3,4 and CFO5 respectively. However, by using the Meservey-Tedrow technique, positive P CFO \nvalues (ranging from +6% to 26% depending on film growth condi tions) have been obtained in \nCFO-based spin filters6. Theoretical calculations of spin-depe ndent electronic structure of spinels \nindicate that the lowest ener gy conduction band is spin-down7, thus predicting a P FI < 0. It thus \nfollows that, in spite of its relevance for further progress in spintronics, no definitive determination \nof P FI is yet available for spinel-based spin filters. \nThe Point Contact Andreev Reflection (PCAR) t echnique has been introduced as a tool to \nmeasure the spin polarization of carriers in ferromagnetic materials8. In the case of a normal \nmetal/superconductor junction, an incoming electron from the normal metal with energy smaller \nthan the superconducting gap cannot enter into the superconducting (S C) electrode and is reflected \nas a hole in the normal metal, simultaneously adding two electrons (a Cooper pair) to the condensate in the superconductor. This process, known as Andreev reflection (AR), causes an \nincrease of the conductance around zero bias G(V≈0) compared to the conductance at voltages \nG(V) well above the superconducting gap ( ∆/e) by a factor of two. Since the reflected hole is \ncreated in the density of states with opposite sp in than the incoming electron, the AR process is \npartially suppressed when the respective densities of states are not equal, as in the case of a ferromagnetic metal/superconductor (FM/ SC) interface. In particular, in fully spin-polarized metals, \nall carriers have the same spin orientation and the AR should be totally suppressed because a hole \ncannot be created in the opposite de nsity of states. Thus, the absolute value ( but not the sign) of \ntransport spin polarization │P│ of a ferromagnetic material can be inferred from the grade by \nwhich AR is suppressed in a measured conductance spectra8,9. It has been recently predicted10 that \nthe insertion of a spin-filtering barrier to form a SC/FI/M structure should lead to the subsequent \nmodification of the AR process by spin-selective tunnelling across the ferromagnetic insulator. \nIn this paper we report PCAR spectrosc opy experiments realized by pressing a \nsuperconducting Nb tip on a spin filter constituted by a CoFe 2O4/SrRuO 3 (CFO/SRO) thin film \nbilayer. We show that AR occurs across the ferrimagnetic CFO tunnel barrier of the \nNb/CoFe 2O4/SrRuO 3 (SC/FI/FM) structure demonstrating spin-preserved tunnelling through the \nCFO barrier. Analysis of the experimental co nductance data by means of the modified BTK \nmodel11 allows determining the polarization P of th e CFO/SRO bilayer. PCAR experiments have \nalso been conducted on the bare SRO layers, and we obtained P ≈ 42(1)%. These results indicate a \nvery modest filtering efficiency of the CFO barriers (< +13%). Implications of these findings are discussed. \n \n \nII. SAMPLE PREPARATION AND PRE-CHARACTERIZATION \n \nCFO/SRO bilayers have been deposited by RF magnetron sputtering from stoichiometric \ntargets, on a single crystalline (111)SrTiO 3 substrate. The bottom electrode is a 25 nm thick \nferromagnetic (T Curie ≈ 120 K) and metallic ( ρ ≈ 200 μΩ •cm at 10K) SRO epitaxial film, deposited \nat ≈0.5 nm/min in a mixed atmosphere of argon and oxygen (ratio 3:2), with a total pressure of 100 \nmTorr; the substrate temperature was 725 ºC. The top layer is a 3 nm thick epitaxial CFO film, \ngrown at ≈ 0.1nm/min with an Ar/O 2 ratio of 10:1, and a total pre ssure of 250 mTorr at 500 °C. The \nroughness of the SRO and the CFO layer are lower than 0.3 nm. Th e saturation magnetization of a \nbare 3 nm thick CFO film12, measured by using a Quantum Design’s SQUID, was ≈524 emu/cm3. \nThis value is somewhat larger than the bulk valu e. Similar enhancement ha d been reported earlier \nfor ultrathin films of other spinel oxides13 and attributed to partial ca tion inversion of the spinel \nstructure and subsequent electronic and magnetic ordering modifications. The bilayer structure was characterized by Conductive-Atomic Force Microscope (C-AFM), \nwith a Nanotec Cervantes AFM, using Nanosensors tips (CDT-NCHR). Electrical measurements \nwere performed in 2-point electr ical configuration: the C-AFM probe was grounded, while the SRO \nbottom electrode was positively biased. C-AFM m easurements were conducted in dry nitrogen \natmosphere, using a feedback-force of ≈850 nN and an applied voltage of 800 mV. \nIn Fig. 1(a) we show a resistance map measured by C-AFM: at each point of the surface the \nsystem records the current between tip and sample under constant bias voltage. This image indicates \na very high electrical hom ogeneity in a large area (3 μm x 3 μ m). The simultaneously recorded \ntopographic maps (not shown) confirmed the absence of particular defects and an extremely smooth \nsurface (rms < 0.3 nm). The histogram of the resist ance values, shown the Fig. 1(a)(inset) indicates \na narrow distribution of log R (half-width at half-maximum ≈ 3.5%), centred around ≈8.25 \n(≈178 MΩ). As the tunnelling current depends exponentia lly on the barrier thic kness, the data in \nFig. 1(a) signals also an extremely sm all variation of the barrier thickness14. Subsequent \nmeasurements were repeated in the same area; neither the re sistance map nor the corresponding \ntopographic image evidenced significant changes in the surface properties other than a minor \nreduction of the overall resistance likely associat ed to residual surface contamination removal. No \ntraces of indentation or scratching could be detected. Tests re peated at different locations on the \nsample surface yield very similar results thus co nfirming the homogeneity of the surface properties. \nThe Current-Voltage ( I-V) curves measured at different points on the surface (one example \nis shown in Fig. 1b) show clear characteristics of tunnelling behaviour. The asymmetry visible in \nI-V curves is due to the asymmetric contact conf iguration for forward and reverse biasing of the \ntunnel structure and it is of no re levance for PCAR experiments, performed in the subgap region, to \nbe described in the following. \n \n \nIII. POINT CONTACT SPECTROSCOPY \n \nPoint contact junctions were formed by pushing a soft Niobium tip, obtained by mechanical \ncutting and chemical etching a Nb wire (diameter ≈ 0.2 mm), onto the CFO surface of the \nCFO/SRO bilayer. The experimental setup utilized for the meas urements has been described \nelsewhere15. Due to the softness of Nb and to the hard ness of the CFO surface, we do not expect \nsignificant scratching or penetration of the CFO film by the tip. The e ffective electric contact radius d can be estimated by using the approximation16 R = 4ρ l / 3πd2 + ρ / 2d and employing the values of \nthe resistance R measured at high bias ( ≈200-300 Ohm), the resistivity of equivalent SRO films \n(ρ ≈ 200 μΩcm) at 4 K and a mean free path l of about 15 Å, estimated from ρ17. It turns out that \nthe estimated contact size d is of about 45-60 Å, implying that although the contacts are larger than \nthe mean free path they are stil l in an intermediate regime d ≈ l. \nWe have recorded conductance curves at low temperature (T = 4.2 K) at different positions \non the film and with different pressures between tip and sample. In Fig. 2 we show representative \nconductance spectra of two different types of junc tions we measured: Junctions 1 and 2. Data have \nbeen normalized by using the background conductance estimated at large voltage (V >> ∆/e) \nregions, where Δ is the superconducting gap of Nb ( Δ ≈ 1.5 meV). Both spectra display a clear \nincrease of the conductance around zero bias suggesting that the tr ansport is mainly due to the \nAndreev reflection process. More over, Andreev reflection, presumab ly spin-filter weighted, takes \nplace across the ferrimagnetic barrier as predicted in Re f. [10]. From data in Fig. 2 it is clear that the \nnormalized conductance G(≈0 V) < 2 thus indicating some suppr ession of AR as expected for a \nspin-polarized electron tunnelling. \nIt is noteworthy in Fig. 2 that the features in the conductanc e spectrum appear at energies \nthat are sensibly higher than what is expected for Nb. This observation is commonly attributed to \nthe presence of a spread resistance R S arising from the resistance of the sample between the junction \nand one of the measuring contacts18. The effect of R S is to shift the coherence peaks from V ≈ Δ to \nlarger voltages and subsequently changing the G (0) values. Following Woods et al.18 the spread \nresistance is included in our modelling of the PCAR curves by considering two contributions to the \nmeasured voltage19: \n () ()()I V I V I VS PC + = (1)\nand so the measured conductance G(V) will be fitted by using \n \n()1−\n⎟\n⎠⎞⎜\n⎝⎛+ = =dIdV\ndIdV\ndVdIV GS PC (2)\n V PC and VS are the voltage drops at the FM/FI/SC junction and at R S, respectively. To \ncalculate VPC(I) we used the modified BTK model11 considering an effective P spin-polarized \ncurrent coming out from the spin-filter. For simplicity, a ballistic regime ( l >> d ) of transport across \nthe junction will be assumed. Its application to the present case, where l ≈ d, at first sight may seem \nproblematic. However, as shown in Ref. [18], th e potential errors introdu ced by applying ballistic formulas to diffusive contacts, as far as Z is not too small, have a negligible impact on the extracted \nP values and the uncertainty is translated into the barrier transparency Z. The other relevant fitting \nparameters are: the superconducting gap Δ and a smearing coefficient ( Γ) that allows modelling of \nthe broadening of the gap edges and in elastic scattering at the interfaces20. The local temperature \nhas been assumed to be the same like measured at the sample holder, which is immersed in a liquid \nHe bath. \nThe spread resistance R S was determined beforehand by setting a gap value Δ equal to the \nBCS value for Nb, thus eliminating a possible degeneracy in Δ and R S. Consequently, Δ cannot be \nconsidered a completely free parameter, but it is set close to the BCS value. In the subsequent \nfitting step, R S was kept fixed to the determined value and P, Z, Δ and Γ were allowed to vary. The \nsolid line in Fig. 2a is the result of the optimal fit of G(V) using P ≈ 39%, Z ≈ 0.13, Γ ≈ 0.22 meV, \nΔ ≈ 1.50 meV and R S/RPC ≈ 0.5. The low interface barrier transparency Z used to fit our \nexperimental data indicates that our measurem ents are not significantly affected by a possible \ndependence of the values of P on Z21-23. The quality of the fit and its robustness on variations of P \ncan be better appreciated in Fig. 2c where we s how the low-voltage zoom of data in Fig. 2a. In \nFig. 2c we also include fits obtained by fixing Δ (1.5 meV) and P to be larger (45%) or smaller \n(35%) than P from optimal fit of Fig. 2a and allowing Z, Γ to vary. We notice in Fig. 2c that fixing \nlarger (smaller) P values lead to smaller (larger) Γ values as predicted in Ref. [22], just illustrating \nthat moderate inelasti c interface scattering ( Γ) decreases the AR probability. Thus the Γ and P \nparameters in the AR spectra mix together a nd distinction among both e ffects is challenging20. In \nspite of this, it is clear from these data that P J1 should be in the 45-35 % range for Junction 1. \nTo treat a possible degeneracy of fits to PCAR spectra, it has been proposed25 to perform fits \nof (Δ, Z, Γ ) for different fixed P trial, and check the resulting sum of the squared deviations χ2 of the \nfits as function of P trial. In Fig. 3, results of such analysis are shown for the data of Junction 1. A \nclearly defined minimum of χ2 at about 39(1)% is found, which falls well within the previously \ndetermined range of values, thus we conclude that P J1 ≈ 39(1)%. However, for the data of Junction \n2 (Fig. 2b, which will be discussed below), such pr ocedure proved to be ill-d efined when using the \nproximity reduced gap, due to the larger number of fit parameters. \nIt has been demonstrated25 that an error in the normalization of the conductance spectra can \nlead to false minima in χ2 vs. P trial. Therefore the χ2 analysis was repeated using a deliberately \nwrong normalization of the data in Fig 2a (divid ed by an additional factor n = 1.001). Still, χ2 \nreproduces the minimum but values of χ2 are considerably enlarged. For normalization just slightly smaller than 1 (factor 0.999) reas onable fits to the data are no longer possible and values of χ2 \nincrease drastically, demonstrating the validity of the original normalization. \nIn Fig. 2b we show an example of a G(V) curve measured in some different contacts \n(Junction 2). Data in Fig. 2b displa y characteristic features of AR reflection but also the occurrence \nof proximity effects and subsequent smaller gap formation ( ΔP)11,24. As CFO is a ferromagnetic \ninsulator it may be supposed that a gap reduction may occur in some part of th e tip, likely due to the \nstray field of the ferrimagnetic barrier. The mo dified BTK model has b een worked out including \nexplicitly two gaps ( Δ and ΔP)11. We used the corresponding expressi ons to fit the da ta in Fig. 2b. \nThe solid line is the result of the optimal fit of G(V) using, Δ ≈ 1.48 meV, ΔP ≈ 0.99 meV, \nΓ ≈ 0.0 meV and R S/RPC ≈ 0.36. The values Z ( ≈0.20) and P (31%) determined for this contact are \nquite similar to those extracted from data of Junc tion 1 (Fig. 2a). In the zoom of the low-voltage \nregion (Fig. 2d) we include the f its obtained by fixing P to larger (35%) or smaller (25%) values \nwhile keeping Δ and Δ P constant. The small Γ values obtained from the best fits ( Γ ≈ 0.1), most \nprobably result from the fact that the broadening of the spectrum is captured during the fit by the \npresence of the two gaps required to fit the di ps. Although the available data do not allow to \ndisentangle both contributions, it is clear from these data that P ≈ 31(3)% is a robust result for \nJunction 2. At this point it is wort h to recall that dips in PCAR c onductance spectra, as they appear \nin the data of Junction 2 have also been explai ned as arising from the superconductor reaching its \ncritical current and thus adding some normal conducting state finite resistance26. Although the \norigin of the dips in the conductance spectr a cannot be decided with certainty, a smaller \nsuperconducting energy gap is in princi ple, consistent with the observation of critical current effects \nin Junction 2. \nTherefore, it follows that Andreev reflecti on has been observed across a ferrimagnetic tunnel \nbarrier (Fig. 4). From the analysis of the PCAR we infer that representative values of the effective \npolarization of our CFO/SRO spin f ilter (3 nm thick) are of about P ≈ 31(3)% and 39(1)% as \ndetermined for Junctions 1 and 2 respectively. To progress further and to extract the spin-filter \nefficiency of the CFO barrier re quires the knowledge of the spin-p olarization of electrons emitted \nfrom the SrRuO 3 electrode (P SRO). \nSignificantly different P SRO results have been reported: Worledge and Geballe27 reported \nMeservey-Tedrow type measurements of tunnel j unctions having SRO as electrode, and determined \nPSRO = -9%; Nadgorny et al.23, using PCAR with Sn tips, inferred a much larger value \n(│PSRO│≈ 53%). Although this strong discrepancy is not fully understood23,27, consensus exist that PSRO < 0. Negative spin-polarization arises from the difference of Fermi velocities for spin-up and \nspin-down electrons emerging from SRO rather than the density of states at the Fermi level (which \nis practically identical)27. \nTo obtain a more reliable basis for assessing the PCAR results from the CFO/SRO structure, \nadditional experiments were perfor med on a SRO film equivalent to the bottom electrodes used in \nthe bilayers. A representative PCAR spectrum of the SRO film is shown in Fig. 5. The polarization \nof SRO was again determined by fitting (solid lin e) to the modified BTK model and resulted in \nPSRO = 42%, somewhat lower but still in agreement w ith previously reported PCAR results within \ntheir variations23. In the inset of Fig. 5, the resulting χ2 for fits with various P trial to the SRO data \nwith a clear minimum at P SRO = 41.5(1)% is shown. \n \n \nIV. DISCUSSION \n \nWe are now in position to compare the effective polarization measured in the spin filter and \nthe bare SRO electrode. One first notes that the values of P extracted from both junctions are \nsmaller than that of the bare SRO electrode. Due to the fact that P SRO is recognized to be negative \n(PSRO < 0), this observation implies that the spin pol arization of the CFO barrier must be positive \n(PCFO > 0). Notice that although PCAR experiments of a single ferromagnetic layer do not allow to \nextract the sign of the spin polar ization, this is possible in th e present structure, where two \nferromagnetic layers are involved, if the sign of the sp in polarizations of one of the layers (SRO in \nthis case) is known. \nIt has been theoretically predic ted that the lowest energy barrie r in the exchange-split gap of \nCFO is the spin-down7; in such circumstances it could be expected that the spin-down channel is \ndominating (P FI < 0) the AR (Fig. 4), and thus one could anticipate an effective P in CFO/SRO \nbilayers larger than in the PCAR of the bare SRO electrode. Experimentally this is not the case as \nPJ1 and P J2 are found to be smaller than P SRO. We will discuss below on this discrepancy. \n We next consider the values of the spin filtering efficiency of the CFO layers. The two \nextracted values of P J1, J2 (≈ 39(1)% and ≈ 31(3)%) represent two distinct situations. Whereas the \nsecond value (Junction 2) would indicat e a substantial spin filtering of CFO, this is not so for data \nfrom Junction 1, where a rather small different with the bare SRO electrode is observed. \nWe can define the spin-filter efficiency of the CFO barrier P CFO by28: \nCFO SROCFO SRO\nP P 1P PP++= (3)\nwhere P is the effective spin polarization measured in the PCAR experiment. Using P SRO = -\n42%, the two solutions of Eq. (3) are P CFO ≈ +4% and +73% for P J1 and P CFO ≈ +13% and +67% for \nPJ2. Among these two sets of possible values, which phys ically arise due to th e fact that PCAR can \nnot determine the sign of the meas ured spin polarization, the larges t pair of values (73% and 67%) \nare most likely inappropriate as they could lead to spin filtering efficiencies for CFO much larger \nthan reported values5,6. On the other hand, we notice that P CFO ≈ +13%, as determined for Junction \n2, is within the range of Meservey-Tedrow results (+6% ≤ PCFO ≤ +26%, depending on the growth \nconditions)6 obtained on CFO barriers of nominally equal thickness (d ≈ 3 nm). Notably, the sign of \nthe effect agrees with the Meservey-Tedrow experi ments, which is, however, different from the one \ndetermined in tunnel magneto resistance measurements using Al 2O3 tunnel barriers. On the other \nhand data for Junction 1 appears to indicate a very marginal spin filtering effect. This could be \nrelated to the fact that the CFO barrier is locally suppressed either by a mechanical effect associated \nto the Nb-tip pressure (although, as mentioned a bove, scratching effects have not been observed \nwhen using the harder C-AFM tip) or a locally poorer homogeneity of the insulating barrier. \nAlthough the CFO layers appears homogenous and r obust in CAFM measurem ents, the possibility \nthat the Nb tip penetrates the CFO layer either through pinholes or completely when crushed onto \nthe sample cannot be excluded. Under such circumstances, it turns out that P J2 and thus P CFO ≈ \n+13% constitutes the most represen tative value of the spin filtering efficiency of the present CFO \nbarriers. It could be argued that a similar result would be obtained if the CFO barrier does not spin-\nfilter at all but only contributes to depolarize the electron current from SRO. However, the experimental observation of spin-filt ering in CFO in tunnel structures\n5,6 does not support this view. \nBefore concluding we would like to comment on the positive sign observed for P CFO. As \nmentioned, this observation is oppos ite to theoretical predictions7, which are based on the electronic \nconfiguration of an ideally inverse spinel structure of CFO where all Co2+ occupy the octahedral \nsites of the unit cell and the Fe3+ equally populate the octahedral and tetrahedral sites. However, \nmagnetization data of nanometric thin films13,29 of various spinels, including CoFe 2O4 and NiFe 2O4 \nhave provided conclusive evidence that the ca tionic distribution in nanometric thin films may \nlargely differ from their bulk c ounterparts. As shown by the calc ulations of Szotek et al.7, the \ninsulating gap of CFO closes by some 75% when th e cationic distribution is not that of the ideal \ninverse spinel structure but a nor mal one. It thus follows that th e tunnel transport may be overcome by other non-spin preserving transp ort channels and therefore a redu ced spin filtering efficiency \ncould be anticipated for cationically -disordered films. To what extent the change of sign of the spin \nfiltering efficiency is related to the same effect or not is not definitely settled. We also notice that \nthe spin filtering efficiency should not be simply related to the ex change splitting of the insulator \nbut the symmetry of the relevant wave functions may also play a role. \n \n \nIV. SUMMARY \n \nIn summary we have shown evidence that Andreev reflection occurs at ferro(ferri)magnetic \ntunnel barriers. Data collected us ing point contact spect roscopy allowed to estimate the effective \nspin-polarization of a current th rough the interface of a spin-fil ter and a superconducting tip. It \nturned out that a possible spin filtering effect of the spinel oxide CoFe 2O4 tunnel barrier is limited to \nabout +13% with the accuracy of these measurements. Observati on of a positive spin filtering \nefficiency is unexpected and may suggest the rele vance of spin-dependent orbital symmetry effects \non the tunnel probability in spin filters. To th e best of our knowledge this issue has not been \ntheoretically addressed yet. \n \n \nACKNOWLEDGEMENTS \n \nFinancial support from the Ministerio de Ci encia e Innovación of th e Spanish Government \nProjects (MAT2008-06761-C03 and NANOSELECT CSD2007-00041) and from the European \nUnion [ProjectMaCoMuFi (FP6-03321) and FEDER and Marie Curie IEF Project SemiSpinNano] \nis acknowledged. S.P. thanks B.L. Gallagher and C. J. Mellor for hosting the final stage of this research. \n REFERENCES \n1 J. S. Moodera, X. Hao, G. A. Gibson, and R. Meservey, Phys. Rev. Lett. 61, 637 (1988). \n2 M. Gajek, M. Bibes, A. Barthélémy, K. Bouzehoua ne, S. Fusil, M. Varela, J. Fontcuberta, and \nA. Fert, Phys. Rev. B 72, 020406(R) (2005). \n3 U. Lüders, M. Bibes, K. Bouzehouane, E. Jacquet, J.-P. Contour, S. Fusil, J.-F. Bobo, J. \nFontcuberta, A. Barthélémy, a nd A. Fert, Appl. Phys. Lett. 88, 082505 (2006). \n4 U. Lüders, A. Barthélémy, M. Bibes, K. Bouzehoua ne, S. Fusil, E. Jacquet, J.-P. Contour, J.-F. \nBobo, J. Fontcuberta, and A. Fert, Adv. Mater. 18, 1733 (2006). \n5 A. V. Ramos, M.-J. Guittet, J.-B. Moussy, R. Ma ttana, C. Deranlot, F. Petroff, and C. Gatel, \nAppl. Phys. Lett. 91, 122107 (2007). \n6 A. V. Ramos, T. S. Santos, G. X. Miao, M.-J. Guittet, J.-B. Moussy, and J. S. Moodera, Phys. \nRev. B 78, 180402(R) (2008). \n7 Z. Szotek, W. M. Temmerman, D. Ködderitzsch, A. Svane, L. Petit, and H. Winter, Phys. Rev. \nB 74, 174431 (2006). \n8 R. J. Soulen Jr., J. M. Byers, M. S. Osofs ky, B. Nadgorny, S. F. Cheng T. Ambrose, P. R. \nBroussard, C. T. Tanaka, J. Nowak, J. S. Moodera, A. Barry, and J. M. D. Coey, Science 282, \n85 (1998). \n9 S. K. Upadhyay, A. Palanisami, R. N. Louie, and R. A. Buhrman, Phys. Rev. Lett. 81, 3247 \n(1998) . \n10 S. Kashiwaya, Y. Tanaka, N. Yoshida, and M. R. Beasley, Phys. Rev. B 60, 3572 (1999). \n11 G. J. Strijkers, Y. Ji, F. Y. Yang, C. L. Chien, and J. M. Byers, Phys. Rev. B 63, 104510 (2001). \n12 F. Rigato et al. unpublished \n13 U. Lüders, M. Bibes, J.-F. Bobo, M. Cantoni, R. Bertacco, and J.Fontcuberta, Phys. Rev. B 71, \n134419 (2005); F. Rigato, S. Estradé, J. Arbiol, F. Peiró, U. L üders, X. Martí, F. Sánchez, J. \nFontcuberta, Mater. Sci. Eng. B 144, 43 (2007). \n14 V Da Costa, M. Romeo, and F. Bardou, J. Magn. Mag. Mat 258-259, 90 (2003). \n15 F. Giubileo, M. Aprili, F. Bobba, S. Piano, A. Scarfato, and A. M. Cucolo, Phys. Rev. B 72, \n174518 (2005). \n16 B. Nikolić and P. B. Allen, Phys. Rev. B 60, 3963 (1999). \n17 G. Herranz, B. Martínez, J. Fontcuberta, F. Sánchez, C. Ferrater, M. V. García-Cuenca, and M. \nVarela, Phys. Rev. B 67, 174423 (2003 ). \n18 G. T. Woods, R. J. Soulen, Jr., I. Mazin, B. Nadgorny, M. S. Osofsky, J. Sanders, H. Srikanth, \nW. F. Egelhoff, and R. Datla, Phys. Rev. B 70, 054416 (2004) \n19 S. Piano, F. Bobba, F. Giubileo, A. M. Cucolo, M. Gombos, and A. Vecchione, Phys. Rev. B 73, 064514 (2006). \n20 P. Chalsani, S. K. Upadhyay, O. Ozatay, and R. A. Buhrmann, Phys. Rev. B 75, 094417 (2007). \n21 V. Baltz, A. D. Naylor, K. M. Seemann, W. Elder, S. Sheen, K. Westholt, H. Zabel, G. Burnell, \nC. H. Marrows, and B. J. Hickey, J. Phys. Condens. Matter 21, 095701 (2009). \n22 N. Auth, G. Jakob, T. Block, and C. Felser, Phys. Rev. B 68, 024403 (2003). \n23 B. Nadgorny, M. S. Osofsky, D. J. Singh, G. T. Woods , R. J. Soulen Jr., M. K. Lee, S. D. Bu \nand C. B. Eom, Appl. Phys. Lett. 82, 427 (2003). \n24 R. P. Panguluri, K. C. Ku, T. Wojtowicz, X. Liu, J. K. Furdyna, Y. B. Lyanda-Geller, N. \nSamarth, and B. Nadgorny, Phys. Rew. B 72, 054510 (2005). \n25 Y. Bugoslavsky, Y. Miyoshi, S. K. Clowes, W. R. Branford, M. Lake, I. Brown, A. D. Caplin, \nand L. F. Cohen, Phys. Rev. B 71, 104523 (2005). \n26 G. Sheet, S. Mukhopadhyay, and P. Raychaudhuri, Phys. Rev. B 69, 134507 (2004). \n27 D. C. Worledge and T. H. Geballe, Phys. Rev. Lett. 85, 5182 (2000). 28 This equation is derived by considering diffe rent transmission coefficient for the two \nindependent spin channels across the spin-filte ring CFO barrier and a spin polarized current \ncoming from the SRO. \n29 F. Rigato, M. Foerster, J. Fontcuberta, unpublished Figure Captions \n \nFig. 1: (Color online) C-AFM analysis of a CF O (3 nm)/ SRO (25 nm) bilayer on (111)STO: (a) \nresistance map in logarithmic color scale; inset: hi stogram of the resistance values over the surface. \n(b) I-V characteristic, with a schematic of the measurement circuit (inset). \n \nFig. 2: (Color online) Measured conductance spectra: (a) Junction 1; (b) Junction 2; the \ncontinuous line represents the best fits. (c), (d) Zoom around low bias; for Junction 1 and 2 \nrespectively: Comparison of the best fits with simulations achieved forcing P to slightly different \nvalues, demonstrating the accuracy of the obtained polarizations. \n \nFig. 3 Characteristics of fits obt ained for data from Junction 1. χ2 as function of P trial for \nstandard normalization (n = 1) and varied normalization (n = 1.001). \n \nFig. 4: (Color online) Schematic of the density of states vs. energy of a SC-FI-FM (Nb-CFO-\nSRO) structure. The Andreev refl ection (AR) process is illustrate d (solid arrow). An incoming \nelectron from the SRO side is reflected as a hole in the spin-reversed density of states, while a \nCooper pair is added to the supe rconducting condensate. Parallel to AR, electron tunnelling into \nthermally excited states may occur (dashed arrow). For simplicity, only spin-down electrons are \ndepicted and low-lying spin-dow n barrier has been assumed. \n \nFig. 5 (Color online) Measured co nductance spectra of a SRO thin film, equivalent to the base \nelectrodes used in CFO/SRO bilayers; the con tinuous line represents th e best fit. Inset: χ2 as \nfunction of P trial for fits to the data in the main panel. \n \n \n \n \n \n \n Figure 1 \n(b)\n-1,0 -0,5 0,0 0,5 1,0-1001020\nSTO (111)SROCFO\nSTO (111)SROCFO\nSTO (111)SROCFO\n Current (nA)\nVoltage (V)\n(a) 10 \n7 Figure 2 \n \n(c) (a) \n(d) (b) Junction 1 Junction 2 \n-20 -10 0 10 200,951,001,051,101,15\nΔNb=1.48 meV\nΔproximity =0.99 meV\nT=4.2 K\nRseries /RPC=0.36\n Z=0.21\nP=0.305\nΓ=0 meV\nV(mV)dI/dVnormalized\n- 6 - 4 - 2 02460,951,001,051,101,151,20dI/dVnormalized\n \n P=0.305\nΓ=0.0 meV\nZ=0.21\nP=0.35\nΓ=0.1 meV\nZ=0.1 exp. data\n P=0.25\n P=0.305\n P=0.35\nV(mV)P=0.25\nΓ=0.18 meV\nZ=0.16-20 -10 0 10 200,951,001,051,101,15\nΔNb=1.48 meV\nΔproximity =0.99 meV\nT=4.2 K\nRseries /RPC=0.36\n Z=0.21\nP=0.305\nΓ=0 meV\nV(mV)dI/dVnormalized\n-20 -10 0 10 200,951,001,051,101,15\nΔNb=1.5 meV \nT=4.2 K\nRseries /RPC=0.5dI/dVnormalized\n Z=0.135P=0.39\nΓ=0.22 meV\nV(mV)\n- 6 - 4 - 2 02460,951,001,051,101,151,20dI/dVnormalized\n P=0.305\nΓ=0.0 meV\nZ=0.21\nP=0.35\nΓ=0.1 meV\nZ=0.1 exp. data\n P=0.25\n P=0.305\n P=0.35\nV(mV)P=0.25\nΓ=0.18 meV\nZ=0.16\n-10 -5 0 5 101,001,051,101,15\nP=0.39\nΓ=0.22 meV\nZ=0.135dI/dVnormalized\n P=0.45\nΓ=0.0 meV\nZ=0.0P=0.35\nΓ=0.52 meV\nZ=0.19 exp. data\n P=0.35\n P=0.39\n P=0.45\nV(mV)Figure 3 \n0,25 0,30 0,35 0,400,01,0x10-52,0x10-53,0x10-5\n \n n=1\n n=1.001χ2\nPtrial\n \n Figure 4 \n \n \n \n \n Figure 5 \n-60 -40 -20 0 20 40 600,981,001,021,041,061,08\n0,36 0,40 0,449,0x10-61,0x10-51,1x10-51,2x10-5\n dI/dVnormalized\nV(mV)Δ = 1.45 meV\nZ = 0.185\nΓ = 0.27 meV\nP = 0.415\nRS/RPC = 1\nχ2\nPtrial\n \n " }, { "title": "2212.11887v2.Spin_wave_dispersion_of_ultra_low_damping_hematite___α_text__Fe__2_text_O__3___at_GHz_frequencies.pdf", "content": "Spin wave dispersion of ultra-low damping hematite ( \u000b-Fe 2O3) at GHz frequencies\nMohammad Hamdi,1,\u0003Ferdinand Posva,1and Dirk Grundler1, 2,y\n1\u0013Ecole Polytechnique F\u0013 ed\u0013 erale de Lausanne (EPFL), Institute of Materials,\nLaboratory of Nanoscale Magnetic Materials and Magnonics, CH-1015 Lausanne, Switzerland\n2\u0013Ecole Polytechnique F\u0013 ed\u0013 erale de Lausanne (EPFL),\nInstitute of Electrical and Micro Engineering, CH-1015 Lausanne, Switzerland\n(Dated: December 26, 2022)\nLow magnetic damping and high group velocity of spin waves (SWs) or magnons are two crucial\nparameters for functional magnonic devices. Magnonics research on signal processing and wave-\nbased computation at GHz frequencies focussed on the arti\fcial ferrimagnetic garnet Y 3Fe5O12\n(YIG) so far. We report on spin-wave spectroscopy studies performed on the natural mineral\nhematite (\u000b-Fe2O3) which is a canted antiferromagnet. By means of broadband GHz spectroscopy\nand inelastic light scattering, we determine a damping coe\u000ecient of 1 :1\u000210\u00005and magnon group\nvelocities of a few 10 km/s, respectively, at room temperature. Covering a large regime of wave\nvectors up to k\u001924 rad=\u0016m, we \fnd the exchange sti\u000bness length to be relatively short and only\nabout 1 \u0017A. In a small magnetic \feld of 30 mT, the decay length of SWs is estimated to be 1.1\ncm similar to the best YIG. Still, inelastic light scattering provides surprisingly broad and partly\nasymmetric resonance peaks. Their characteristic shape is induced by the large group velocities, low\ndamping and distribution of incident angles inside the laser beam. Our results promote hematite as\nan alternative and sustainable basis for magnonic devices with fast speeds and low losses based on\na stable natural mineral.\nIntroduction. | Spin waves (magnons) are collective\nspin excitations in magnetically ordered materials.\nThey exhibit promising functionalities for information\ntransmission and processing at GHz frequencies [1{3].\nTo realize energy e\u000ecient magnonic circuits [1, 2, 4{6]\nisotropic spin wave (SW) dispersion relations, high group\nvelocities, and low magnetic damping are essential. Until\ntoday, the arti\fcial garnet Y 3Fe5O12(YIG) [7] played a\nkey role for the exploration of magnonics functionalities\n[8]. Already in 1961, M. Sparks et al. coined the term\nthat YIG was to ferromagnetic resonance research what\nthe fruit \ry was to genetics research [9]. This was\nparticularly true for high-quality YIG grown by liquid\nphase epitaxy on the wafer scale [8, 10]. However,\nin a ferrimagnetic material like YIG, magnon bands\nin the regime of small wave vectors, k, and low GHz\nfrequencies are inherently anisotropic due to the dipolar\ninteraction between spins. To overcome this, a lot of\ne\u000bort has been put into the development of microwave-\nto-magnon transducers which allow for the excitation of\nexchange dominated SWs with isotropic properties at\nhigh frequencies [11{13]. The transducers su\u000ber however\nfrom typically a narrow bandwidth or require an applied\nmagnetic \feld in contrast to conventional transmission\nlines and coplanar waveguides (CPWs).\nIn antiferromagnetic (AFM) materials, exchange\ninteraction dominates the dispersion relation already\nat small wave vectors k. The dipolar interactions are\nvirtually absent due to net zero magnetization. Still,\nSWs can propagate with high group velocities. Values\nsimilar to thick YIG [8, 14] and as high as 30 km/s\nhave been reported [15{18]. However, the challenge\nwith most AFMs is their net zero magnetization and\nsub-THz frequencies which make on-chip integrationhard due to lack of e\u000ecient CPWs and THz sources\n(THz gap) [19, 20]. Recently, the natural mineral and\ncanted antiferromagnet hematite ( \u000b-Fe2O3) [21] gained\nparticular attention for magnonics [22, 23] after the\nobservation of long distance spin transport [24, 25]\nand enhanced spin pumping [26]. It is known that,\ndue to extremely low anisotropy in the basal plane\n[21, 27, 28], in the canted phase [Fig. 1(a) and (b)]\none branch of the magnon modes resides at around 10\nGHz at small k. Depending on the purity of hematite\ncrystals, a damping coe\u000ecient as low as 7 :8\u000210\u00006was\nreported for the magnetic resonance [29]. The hematite's\n\fnite net magnetization and strikingly small damping\nof below 10\u00005make it hence suitable for magnonic\napplications. They allow for inductive coupling to\nCPWs and long-distance SW transport, respectively.\nHowever, there is no experiment reporting a measured\nSW dispersion for kvalues accessible by CPWs with\nintegrated microwave-to-magnon transducers [11{13].\nThe dispersion measured over a large wave vector regime\nis of fundamental importance as it allows one to quantify\nthe exchange sti\u000bness length lewith large precision. le\nis the key parameter to estimate the maximum possible\nspin-wave velocity of hematite in the GHz frequency\nregime.\nHere, we study the magnon band structure of bulk\nhematite at di\u000berent wave vectors kby means of broad-\nband microwave spectroscopy and k-resolved inelastic\nBrillouin light scattering (BLS) (Fig. 1). Using a\nCPW in \rip-chip con\fguration we extract a magnetic\ndamping parameter of 1 :12\u000210\u00005which is similar to\nthe best YIG reported in Ref. 30. Still, the measured\nspectra with BLS show broad linewidths. Our modelling\nsubstantiates that the linewidth is explained by thearXiv:2212.11887v2 [cond-mat.mtrl-sci] 23 Dec 20222\n(b)\n(a)\n(c)\n(d)\n(e)\nFIG. 1. Hexagonal unit cell of the crystal structure of\nhematite from (a) the side and (b) top view. The cyan spheres\nand red arrows indicate the Fe atoms and the spins associated\nwith them, respectively (oxygen atoms are not shown). We\ndepict the canted antiferromagnetic state above the Morin\ntemperature for which the sublattice spins lie in the c-plane\nalong thea-axis with a small canting. They give rise to sub-\nlattice magnetization vectors M1andM2. (c) Sketch of the\nlow frequency quasi-ferromagnetic mode where a small mag-\nnetization (green arrow), m, precesses elliptically around the\napplied \feld (black arrow), H. (d) Schematics of the \rip-\nchip VNA measurement. The sample (gray disk) is placed\non a CPW. The static magnetic \feld, H, is applied in the\na-plane and with an angle, \u0012H, to the normal of the c-axis\nof the crystal. The rf magnetic \feld (orange double-headed\narrow), hrf, of the CPW is parallel to the c-axis. (e) Sketch\nof the BLS con\fguration. The magnetic \feld, H, is applied\nperpendicular to c-axis in the a-plane. The laser light (green)\nforms a Gaussian beam and is focused on the surface of the\nsample (gray). The cone angle of the objective lens is \u0012. The\nincident laser light with an incidence angle, 'is scattered by\nmagnons. We measure in back-scattering geometry (dashed\ngreen arrow).\nSW dispersion relation, the large SW velocity and the\nGaussian pro\fle of the laser used for inelastic light\nscattering. The data substantiate a high group velocity\nof 10 km/s for k= 2:5 rad/\u0016m which are excited\neasily by a micron-sized CPW [23, 31]. In an applied\n\feld of 90 mT, the velocity increases to 16 km/s near\nk= 5 rad/\u0016m and levels o\u000b to 23.3 km/s for k\u001525rad/\u0016m. The latter value has routinely been realized\nby transducers. Our \fndings substantiate hematite as\na very promising candidate for a sustainable future of\nmagnonics as its growth avoids the lead-based synthesis\nroute used for high-quality YIG [7, 30].\nProperties of hematite. | We \frst brie\ry review the\nrelevant magnetic properties of \u000b-Fe2O3. Hematite is\nthe stable end product of oxidation of magnetite [32] and\nknown for its great abundance as well as stability in an\naqueous environment [33]. It is an insulating antiferro-\nmagnet (AFM) with a corundum crystal structure. The\narrangement of the magnetic atoms of Fe in the crystal\nis shown in Fig. 1(a) [34]. At room temperature and\nabove the Morin transition temperature of TM= 262 K\nthe Fe+3magnetic moments lie in the c-plane due to an\neasy plane anisotropy, HA, and stack antiferromagneti-\ncally along the c-axis [Fig. 1(a)] [21, 28]. Within the\nc-plane, there is a weak 6-fold anisotropy around the c-\naxis,Ha, which favors the magnetic moments to align\nwith thea-axes [Fig. 1(b)]. The magnetic moments of\nthe two AFM sublattices are slightly canted away from\nthea-axis by the Dzyaloshinskii-Moriya (DM) interac-\ntion (Fig. 1(b)), resulting in a weak magnetic moment,\nm, perpendicular to both a- andc- axes at equilibrium\n[21, 28]. The magnetization amounts to about 2 kA/m\n[32]. The magnetization dynamics of hematite in this\nweak ferromagnetic state o\u000bers two modes namely the\nquasi-ferromagnetic mode (qFM or low frequency mode)\nand quasi-antiferromagnetic mode (qAFM or high fre-\nquency mode) [21, 27, 28].\nHere, we explore the qFM mode schematically depicted\nin Fig. 1 (c). The canted AFM sublattice magnetization\nvectors precess elliptically around their equilibrium direc-\ntion. This results in an elliptical precession of the weak\nmagnetic moment, m(green arrow in Fig. 1 (c)), around\nthe applied \feld, H[21, 27, 28]. The frequency of the\nqFM mode was derived by Pincus [21, 27] according to\nfr= (1)\nj\rj\u00160\n2\u0019p\nHsin\u0018(Hsin\u0018+HD) + 2HE(Ha+HME);\nwhere,\r,\u00160,HE,HDandHMEis the electron gyro-\nmagnetic ratio, vacuum permeability, exchange, DM and\nspontaneous magnetoelastic e\u000bective \feld, respectively.\n\u0018=\u0019=2\u0000\u0012His the polar angle between Hand thec-\naxis (z-direction). We de\fne \u0012Hin Fig. 1(d). Fink [35]\nderived the dynamic susceptibility \u001fzz(f;fr) for the qFM\nmode. The real and imaginary parts read\nRe [\u001fzz(f;fr)] =\u0000\nf2\nr\u0000f2\u0001\nf2\nr\n(f2r\u0000f2)2+ \u0001f2f2and (2)\nIm [\u001fzz(f;fr)] =ff2\nr\u0001f\n(f2r\u0000f2)2+ \u0001f2f2; (3)3\nrespectively. Here, fis the frequency of the radiofre-\nquency (rf) magnetic \feld hrf(Fig. 1 (d)). The frequency\nlinewidth, \u0001 f, is related to the magnetic damping pa-\nrameter,\u000b, by\n\u0001f\u00192\u000bHE(\r=2\u0019); (4)\nforH\u001cHE. Unlike ferromagnets and uniaxial antifer-\nromagnets, the resonance line width of a canted antifer-\nromagnet does not depend on frand is governed by the\n\feld-independent exchange frequency. Following Turov\n[15, 16], the SW dispersion of the qFM mode for hematite\nneark= 0 is given by\nfm(k) = (5)\nj\rj\u00160\n2\u0019p\nH(H+HD) + 2HE(Ha+HME+Ak2);\nwhereA=HEl2\neis the dispersion coe\u000ecient and leis the\ne\u000bective magnetic lattice parameter or exchange sti\u000bness\nlength. For Eq. 5, we considered the experimental\ngeometry for k-resolved BLS measurements with an\nangle\u0018=\u0019=2 as described below.\nExperimental techniques. | The broadband microwave\nspectroscopy was conducted at room temperature above\nTM(Fig. 1(d)). The disk-shaped a-plane\u000b-Fe2O3crystal\nhad a diameter of 2.3 mm and thickness of 0.5 mm. Mea-\nsurements were done in \rip-chip con\fguration for which\nthe sample was placed on a CPW with signal (ground)\nline width of 165 \u0016m (295\u0016m). Thea-axis of the crys-\ntal was perpendicular to the disk plane. The c-axis was\nin the plane and perpendicular to the CPW axis. In-\njecting a radio-frequency (rf) current, Iin\nrf, into the CPW\nby port 1 of a vector network analyzer (VNA) induced\nthe dynamic magnetic \feld hrf(orange double-head ar-\nrow). The rf current was collected on the other end of\nthe CPW by port 2 of the VNA ( Iout\nrf). The static mag-\nnetic \feld, H, was applied in the a-plane of the crystal\nin all VNA measurements ( yz-plane in Fig. 1 (d)). For\nthe angle dependent measurements an external \feld of\n90 mT was applied and the angle, \u0012H, was varied in steps\nof \u0001\u0012H= 2\u000e. In case of \feld sweep measurements, the\napplied \feld angle was perpendicular to the c-axis and in\nthea-plane of the crystal ( \u0012H= 0 deg). The \feld ampli-\ntude was varied in steps of \u0001 H= 0:5 mT.\nWave vector-resolved BLS measurements were done for\n\u0012H= 0 deg on a piece of the same crystal in back-\nscattering geometry (Fig. 1 (e)) using a green laser with\nwavelength, \u0015= 532 nm and wave vector k0= 2\u0019=\u0015=\n11:81 rad/\u0016m. The external \feld was applied in the a-\nplane and perpendicular to c-axis. The sample for BLS\nwas irregularly shaped. We ensured that it was tilted\nwith respect to the incident laser beam along the y-axis in\nsuch a way that the laser beam remained in the xz-plane\nformed by the a- andc-axis. Since the penetration depth\nof the green laser in hematite is on the order of 75 nm[36], linear momentum conservation holds only for the in-\nplane component of the transferred wave vectors. There-\nfore, we de\fne the transferred momentum from the light\nto magnons along the c-axis askz= 2(k0sin'+k0\nxcos'),\nwhere'is the angle between the incident beam and the\nnormal to the plane of the sample ( a-plane) [37]. We\nFIG. 2. (a) Color-coded \feld dependent Im( U) parame-\nter measured on the sample as indicated in Fig. 1(d) with\n\u0012H= 0 deg. Solid red circles and wine triangles are the fre-\nquencies extracted by \ftting Eq. 7 on the data. Blue solid\nline is obtained by \ftting Eq. 1 with extracted f1values. (b)\nImaginary and (c) real part of Uexp(red symbols) and U\ft\n(black curves) for 70 mT, respectively. (d) Imaginary and (e)\nreal parts of \u001f1(red curve), \u001f2(blue curve) and their sum\n(gray curve) for 70 mT, respectively.\nassume a Gaussian beam pro\fle giving rise to a Fourier\ntransform of the beam intensity Ias\nI(k0\nx) =ek02\nxw2\n0=2(6)\nwithw0=2\nk0NA.NAis the numerical aperture of the\nlens [38, 39]. The momentum k0\nxhas a projection on\nz-direction due to the focusing of the beam.\nBroadband microwave spectroscopy data. | Field-\ndependent VNA spectra are shown in Fig. 2 (a). We\ndepict the imaginary part of the quantity U(f;H) =\niln [S21(f;H)=S21(f;H = 0)] in a color-coded plot,\nwhereS21is the complex scattering parameter measured4\nby the VNA at a given \feld, H. We identify two branches\nwhich we label f1andf2for positive \felds. For a detailed\nanalysis, we consider that the parameter Ucontains the\nsusceptibility \u001fof the sample [40, 41] and the electro-\nmagnetic response of the rf circuit used in the \rip-chip\nmethod. To account for the di\u000berent contributions, we\nfollow Refs. [40, 41] and \ft the measured Uwith\nU\ft= (7)\nC[1 +\u001f0+\u001f(f;f1;\u0001f1)ei\u001e1+\u001f(f;f2;\u0001f2)ei\u001e2]\nas shown in Fig. 2(b) and (c) (black lines). Cis a\nreal-numbered scaling parameter, \u001f0is a complex-\nnumbered o\u000bset parameter, and \u001eiare phase shift\nadjustments ( i= 1;2). Using Eq. 7, we extract the\nresonance frequencies and linewidths \u0001 fifor the two\nmodes labelled by f1andf2. In Fig. 2 (b) and (c),\nwe display the measured imaginary and real parts of\nUwith red symbols. In Fig. 2 (d) and (e), we show\nthe extracted real and imaginary parts together with\nthe total susceptibilities (gray lines) from which the\ntwo resonant modes are identi\fed. Their frequencies\nextracted for di\u000berent Hare depicted in Fig. 2 (a) with\nsolid red circles and magenta triangles. We focus on\n\felds larger than 50 mT to ensure a saturated state.\nThe blue line in Fig. 2(a) results from \ftting\nEq. 1 to branch f1with\u0018=\u0019=2. From the \ft,\nwe obtain \u00160HE= 1003:61 T,\u00160HD= 2:34 T and\n\u00160(Ha+HME) = 88:64\u0016T. Using Eq. 4 and the\nexperimental value of \u0001 f1= 0:63 GHz for branch f1, we\nextract a damping parameter of \u000b= 1:12\u000210\u00005. This\nvalue is similar to the best value reported for YIG in\nRef. [30]. The branch f2will be discussed later.\nAngle-dependent VNA spectra are shown in Fig.\n3. To enhance the signal to noise ratio, we depict\n0 90 180 270 3603691215182124Fitted Curve\n-3E-3-2E-3-1E-30E-31E-32E-33E-3\nFIG. 3. Color-coded neighbor subtracted angle dependent\nVNA-FMR spectra measured on the sample as indicated in\nFig. 1(d) with \u00160H= 90 mT. Orange curve depicts Eq. 1 by\nextracted material parameters.\n\u0001jS12j=jS12(H+ \u0001H)\u0000S12(H)j. For such data, a\nzero-crossing (highlighted by the red curve) representsthe resonance frequency. The spectra show a 2-fold\nsymmetry as expected for the c-axis being in the plane\nof the applied magnetic \felds. Introducing the extracted\nparameters discussed above, Eq. 1 models well the\nresonance frequency of branch f1as a function of\nangle\u0012H(red line). The angular dependency is hence\nconsistent with the e\u000bective anisotropy \feld extracted\nfrom the \feld-dependent data.\nBLS data. | The BLS spectra for di\u000berent values of\nincident angle, ', are shown in Fig. 4 (a). Black sym-\nbols in Fig. 4 (b) depict the frequency position of the\nmaximum BLS peak as a function of wave vector calcu-\nlated from the incidence angles shown in the legend of\nFig. 4 (a). The red curve in Fig. 4 (b) is obtained by\nconsidering \frst the material parameters obtained from\nVNA measurements and then \ftting Eq. 5 to the black\nsymbols in the same graph. We obtain a dispersion co-\ne\u000ecient of\u00160A= 9:153\u000210\u00006T.\u0016m2which leads to an\nexchange sti\u000bness length of le= 0:955\u0017A. Using Eq. 5\nand the obtained parameters we calculate the SW group\nvelocity,vg, for\u00160H= 90 mT [blue line in Fig. 4 (c)].\nThe velocity vgincreases signi\fcantly with kand lev-\nels o\u000b at 23.3 km/s. Such a high group velocity is the\ndirect result of the strong exchange interaction and at\nthe same time vanishingly small net magnetization. Mi-\ncrostructured CPWs used in magnonics o\u000ber spin-wave\nwave vectors around k= 2:5 rad/\u0016m [23, 31]. For such a\nvaluek, the qFM spin wave in hematite exhibits a group\nvelocity ofvg= 10 km/s similar to SWs in thick YIG and\nabout a factor of 10 larger compared to ultra-thin YIG\n[42]. The decay length of the SWs is given by ld=vg\u001c,\nwhere\u001c= (1=2\u0019\u000bfm(k)) is the relaxation time of the\nSW. For 30 mT applied \feld and k= 2:5 rad/\u0016m we\ncalculatevg= 13:4 km/s and \u001c= 840:6 ns which leads\ntold= 11:3 mm, again similar to thick YIG. This is due\nto low damping and high group velocity of the qFM spin\nwaves in hematite.\nDespite the small damping of the hematite sample\nmeasured by the VNA, the resonance peaks of the qFM\nmode in the BLS spectra show a large width of 10 to 20\nGHz. Furthermore, the BLS peak taken at nearly nor-\nmal incidence of the laser beam (e.g. at '= 0 or 5 deg)\nshows a strong asymmetric shape. As 'increases, the\nintensity of the prominent peak reduces and it becomes\nmore and more symmetric. To understand these observa-\ntions, we calculated the partial density of states (pDOS)\nby considering the magnon dispersion relation and the\nGaussian beam pro\fle according to\npDOS (f;') = (8)\n\u00001\n\u0019Z+1\n\u00001dk0\nxcos'Im\"\nI(k0\nx)\nf\u0000fm(k) +i\u0001f#\n:\nAssuming only fully back re\rected light we have\nk= 2(k0sin'+k0\nxcos'). \u0001fis the resonance broaden-5\n255075100200400600800100012001400160018002000\n 0 deg \n5 deg \n10 deg \n15 deg \n20 deg \n25 deg \n30 deg \n35 deg 40 deg \n45 deg \n50 deg \n55 deg \n60 deg \n65 deg \n70 deg(a)( b)(\nc)(\nd)05101520250510152025\n153045607590 \n \nFIG. 4. (a) BLS spectra for di\u000berent incident angles, ', mea-\nsured as indicated in Fig. 1(e) with \u00160H= 90 mT. (b) Ex-\ntracted peak maxima frequency (black squares) as a function\nof corresponding transferred wave vector, kand \ftted disper-\nsion relation (red curve) by Eq. 5. (c) Group velocity (d)\npartial density of states and obtained from \ftted magnon dis-\npersion at\u00160H= 90 mT. Inset in (d) depicts pDOS for '= 0\ndegree.\ning due to the damping given by Eq. 4. The numerical\naperture of the objective lens was NA = 0:18. The\ncalculated pDOS for di\u000berent incident angles is plotted\nin Fig. 4 (d). Comparing Figs. 4(a) and (c), the\nexperimental BLS peaks are broad because they contain\na certain frequency regime of the band structure which\nis determined by the Gaussian wave vector distribu-\ntion ofk= 2k0sin'+ \u0001k. The central wave vector\niskc= 2k0sin'and the distribution function for\n\u0001k= 2k0\nxcos'is given by I(k0\nx) (Eq. 6). We attribute\nthe broad BLS peaks hence to the high group velocity\nof magnons which leads to a peak width proportional to\nvg\u0002\u0001k. The asymmetry of BLS peaks for small 'can\nbe understood in terms of the Van Hove singularity of\nthe pDOS near k= 0. As'increases, the part of the\nband structure which is relevant for the collected light\nshifts away from k= 0. Consequently, the peak gets\nmore symmetric with '. The peak intensity reduces\nwith'due to the factor cos 'in Eq. 8. We note that\nEq. 8 does not include all possible scattering processes.Still it provides a good qualitative understanding about\nthe contributions giving rise to the characteristics shape\nand signal strength of BLS peaks as a function of '.\nDiscussion. | The branch of mode f2in Fig. 2(a)\nis higher by about \u000ef= 1 GHz than f1at the same\n\feldH. A second branch was reported also in Ref. 43.\nHere the authors studied the qAFM mode at zero \feld\nand assumed a distribution of magnetic domains. We\napplied a large enough magnetic \feld to avoid domains.\nA domain formation can not explain our second branch.\nAnother possibility for f2is a standing wave along the\nthickness of the sample or a SW excited with a discrete\nwave vector coming from the CPW. However, for these\ncases a quantitative estimate based on the magnon\ndispersion obtained with BLS (Eq. 5) led to a frequency\nseparation of much smaller than 1 GHz.\nThe remaining explanation for f2is a nonuniform\nmagnetoelastic \feld in the sample induced when \fxing\nthe sample on the CPW. We attribute the observed\nfrequency splitting to di\u000berent strains in di\u000berent parts\nof the sample that is either in contact with the CPW\nconductors or \roating on the CPW gaps. A di\u000berence of\n\u000eHME= 24\u0016T would account for \u000ef=f2\u0000f1= 1 GHz.\nThe e\u000bect of magnetoelastic interaction by unidirec-\ntional compression, p, in the basal plane of the crystal,\non the qFM mode is given by replacing HMEwith\nH0\nME=HME\u0000Rpcos 2 [44, 45]. Here, R= 287\n\u0016T/bar [45] is a coe\u000ecient determined by the elastic and\nmagnetoelastic parameters of the crystal and =\u0019=2\nis the angle between pandH. Using\u000eHME=Rp\nwe obtainp= 0:084 bar. This value corresponds to a\nforce of 14.6 mN on the parts of the crystal that are in\ncontact with the CPW conductors. The evaluated force\nis one order of magnitude smaller than the weight of our\nsample and indicates the known sensitivity of hematite\ntowards magnetoelastic e\u000bects.\nFrom our experiments, we do not determine Haand\nHMEseparately as they enter Eq. 1 in the same way.\nTo separate the small HafromHMEan angle dependent\nVNA measurement with the magnetic \feld applied\nin thec-plane of the crystal would be required. The\nexisting sample was not suitable for that.\nOur BLS data were acquired on a seven times larger\nwave vector regime compared to the data used in Ref. 23\nwhere, the authors extracted an exchange sti\u000bness length\nleof 1:2\u0017A. From our extended dataset, we evaluate\nthe valuele= 0:96\u0017A. The precise determination of le\nis of key importance as it determines the SW group\nvelocity. Contrary to the ferrimagnetic YIG, dipolar\ne\u000bects are not expected to play an important role in SW\ndispersion of hematite. Hence, hematite thin \flms can\nprovide similarly large spin-wave velocities as reported\nhere for the bulk crystal. As a consequence, they\npotentially outperform YIG thin \flms concerning speed\nand decay lengths at wave vectors which are realized by6\nthe state-of-the-art transducers. Furthermore, hematite\nis based on earth abundant elements and as an end\nproduct of oxidation of magnetite a stable natural\nmineral suggesting sustainable synthesis routes.\nConclusion. | We have measured the magnon dis-\npersion relation of hematite for wave vectors kwhich\nare relevant for timely experiments in magnonics. The\ndamping coe\u000ecient of the studied natural crystal was\n1:1\u000210\u00005at room temperature. This value is only 40 %\nlarger than the best value reported for pure hematite\nand is already as good as the best YIG. The estimated\nspin-wave decay length for k= 2:5 rad/\u0016m is larger than\n1 cm in a small magnetic \feld. In optimized thin \flms,\ncurrent microwave-to-magnon transducers are expected\nto achieve larger group velocities in hematite than in\nYIG. The reported properties suggest that hematite can\nbecome the fruit \ry of sustainable modern magnonics.\nNote added. | While completing the manuscript\nabout our experiments on hematite [22], we became\naware of Ref. 23. The authors measured group velocities\nand the spin-wave dispersion via integrated CPWs for\n1 rad=\u0016m\u0014k\u00143:5 rad=\u0016m. Our BLS measurements\ncover a seven times larger regime of kvalues enabling an\nimproved evaluation of the parameter le. This parameter\nis decisive to estimate the saturation velocity of GHz\nSWs in hematite.\nThe scienti\fc colour maps developed by Crameri et.\nal. [46] is used in this study to prevent visual distortion\nof the data and exclusion of readers with colour-vision\nde\fciencies [47].\nAcknowledgement. | The authors thank SNSF for \f-\nnancial support via grant 177550. We acknowledge dis-\ncussions with A. Mucchietto, M. Mruczkiewicz, S. Niki-\ntov and A. Sadovnikov. The di\u000berent pieces of the\nhematite crystal were provided by J.-P. Ansermet and\nM. Bialek. We thank them for the support and discus-\nsions.\n\u0003mohammad.hamdi@ep\r.ch,\nmohamad.hamdi90@gmail.com\nydirk.grundler@ep\r.ch\n[1] V. V. Kruglyak, S. O. Demokritov, and D. Grundler, J.\nPhys. D. Appl. Phys. 43, 264001 (2010).\n[2] A. Barman, G. Gubbiotti, S. Ladak, A. O. Adey-\neye, M. Krawczyk, J. Grafe, C. Adelmann, S. Coto-\nfana, A. Naeemi, V. I. Vasyuchka, B. Hillebrands,\nS. A. Nikitov, H. Yu, D. Grundler, A. V. Sadovnikov,\nA. A. Grachev, S. E. Sheshukova, J. Y. Duquesne,\nM. Marangolo, G. Csaba, W. Porod, V. E. Demidov,\nS. Urazhdin, S. O. Demokritov, E. Albisetti, D. Petti,\nR. Bertacco, H. Schultheiss, V. V. Kruglyak, V. D.\nPoimanov, S. Sahoo, J. Sinha, H. Yang, M. M unzenberg,T. Moriyama, S. Mizukami, P. Landeros, R. A. Gallardo,\nG. Carlotti, J. V. Kim, R. L. Stamps, R. E. Camley,\nB. Rana, Y. Otani, W. Yu, T. Yu, G. E. Bauer, C. Back,\nG. S. Uhrig, O. V. Dobrovolskiy, B. Budinska, H. Qin,\nS. Van Dijken, A. V. Chumak, A. Khitun, D. E. Nikonov,\nI. A. Young, B. W. Zingsem, and M. Winklhofer, J. Phys.\nCondens. Matter 33, 413001 (2021).\n[3] K. Baumgaertl and D. Grundler\n10.48550/arxiv.2208.10923 (2022).\n[4] V. S. Tkachenko, A. N. Kuchko, M. Dvornik, and V. V.\nKruglyak, Appl. Phys. Lett. 101, 152402 (2012).\n[5] K. Vogt, H. Schultheiss, S. Jain, J. E. Pearson, A. Ho\u000b-\nmann, S. D. Bader, and B. Hillebrands, Appl. Phys. Lett.\n101, 042410 (2012).\n[6] S. Mieszczak, O. Busel, P. Gruszecki, A. N. Kuchko,\nJ. W. K los, and M. Krawczyk, Phys. Rev. Appl. 13,\n054038 (2020), arXiv:2001.11356.\n[7] D. C. Doughty and E. A. D. White, Acta Crystallogr.\n13, 761 (1960).\n[8] A. A. Serga, A. V. Chumak, and B. Hillebrands, J. Phys.\nD. Appl. Phys. 43, 264002 (2010).\n[9] M. Sparks, R. Loudon, and C. Kittel, Phys. Rev. 122,\n791 (1961).\n[10] S. Maendl, I. Stasinopoulos, and D. Grundler, Appl.\nPhys. Lett. 111, 012403 (2017).\n[11] H. Yu, O. D'Allivy Kelly, V. Cros, R. Bernard,\nP. Bortolotti, A. Anane, F. Brandl, F. Heimbach, and\nD. Grundler, Nat. Commun. 7, 11255 (2016).\n[12] K. Baumgaertl, J. Gr afe, P. Che, A. Mucchietto,\nJ. F orster, N. Tr ager, M. Bechtel, M. Weigand,\nG. Sch utz, and D. Grundler, Nano Lett. 20, 7281 (2020).\n[13] P. Che, K. Baumgaertl, A. K\u0013 ukol'ov\u0013 a, C. Dubs, and\nD. Grundler, Nat. Commun. 11, 1445 (2020).\n[14] A. V. Chumak, A. A. Serga, and B. Hillebrands, J. Phys.\nD. Appl. Phys. 50, 244001 (2017).\n[15] E. Turov, J. Exp. Theor. Phys. 9, 890 (1959).\n[16] F. Ke\u000ber, Spin Waves , edited by H. P. J. Wijn (Springer,\n1966) pp. 1{273.\n[17] E. J. Samuelsen and G. Shirane, Phys. status solidi 42,\n241 (1970).\n[18] J. R. Hortensius, D. Afanasiev, M. Matthiesen, R. Leen-\nders, R. Citro, A. V. Kimel, R. V. Mikhaylovskiy, B. A.\nIvanov, and A. D. Caviglia, Nat. Phys. 17, 1001 (2021),\narXiv:2105.05886.\n[19] C. Sirtori, Nature 417, 132 (2002).\n[20] I. S. Osborne, Science 320, 1262 (2008).\n[21] A. H. Morrish, Canted Antiferromagnetism: Hematite\n(World Scienti\fc, 1995).\n[22] M. Hamdi, F. Posva, and D. Grundler, in 67th Annu.\nConf. Magn. Magn. Mater. (MMM 2022) (2022) pp.\nGOE{12.\n[23] H. Wang, R. Yuan, Y. Zhou, Y. Zhang, J. Chen, S. Liu,\nH. Jia, D. Yu, J.-P. Ansermet, C. Song, and H. Yu,\n(2022), arXiv:2211.10989.\n[24] R. Lebrun, A. Ross, S. A. Bender, A. Qaiumzadeh,\nL. Baldrati, J. Cramer, A. Brataas, R. A. Duine, and\nM. Kl aui, Nature 561, 222 (2018).\n[25] R. Lebrun, A. Ross, O. Gomonay, V. Baltz, U. Ebels,\nA. L. Barra, A. Qaiumzadeh, A. Brataas, J. Sinova, and\nM. Kl aui, Nat. Commun. 11, 6332 (2020).\n[26] H. Wang, Y. Xiao, M. Guo, E. Lee-Wong, G. Q. Yan,\nR. Cheng, and C. R. Du, Phys. Rev. Lett. 127, 117202\n(2021).\n[27] P. Pincus, Phys. Rev. Lett. 5, 13 (1960).7\n[28] J. O. Artman, J. C. Murphy, and S. Foner, Phys. Rev.\n138, A912 (1965).\n[29] C. W. Searle and S. T. Wang, J. Appl. Phys. 39, 1025\n(1968).\n[30] M. Shone, Circuits Systems Signal Process 4, 89 (1985).\n[31] S. Watanabe, V. S. Bhat, K. Baumgaertl, M. Hamdi, and\nD. Grundler, Sci. Adv. 7, eabg3771 (2021).\n[32] B. M. Moskowitz, M. Jackson, and V. Chandler, in Trea-\ntise Geophys. Resour. Near-Surface Earth , Vol. 11, edited\nby G. Schubert (Elsevier, 2015) 2nd ed., Chap. 11.05, pp.\n139{174.\n[33] B. Iandolo, B. Wickman, I. Zori\u0013 c, and A. Hellman, J.\nMater. Chem. A 3, 16896 (2015).\n[34] A. H. Hill, F. Jiao, P. G. Bruce, A. Harrison, W. Kock-\nelmann, and C. Ritter, Chem. Mater. 20, 4891 (2008).\n[35] H. J. Fink, Phys. Rev. 133, A1322 (1964).\n[36] M. R. Querry, Optical Constants , Tech. Rep. (Missouri\nUniversity, 1984).\n[37] T. Sebastian, K. Schultheiss, B. Obry, B. Hillebrands,\nand H. Schultheiss, Front. Phys. 3, 35 (2015).\n[38] L. Novotny and B. Hecht, Principles of Nano-Optics\n(Cambridge University Press, 2012) p. 584.[39] U. Levy, N. Davidson, and Y. Silberberg, Adv. Opt. Pho-\ntonics 11, 828 (2019).\n[40] S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L.\nSchneider, P. Kabos, T. J. Silva, and J. P. Nibarger, J.\nAppl. Phys. 99, 093909 (2006).\n[41] C. Bilzer, T. Devolder, P. Crozat, C. Chappert, S. Car-\ndoso, and P. P. Freitas, J. Appl. Phys. 101, 074505\n(2007).\n[42] H. Yu, O. d. Kelly, V. Cros, R. Bernard, P. Bortolotti,\nA. Anane, F. Brandl, R. Huber, I. Stasinopoulos, and\nD. Grundler, Sci. Rep. 4, 6848 (2014).\n[43] M. Bialek, J. Zhang, H. Yu, and J. P. Ansermet, Appl.\nPhys. Lett. 121, 032401 (2022).\n[44] R. Z. Levitin, A. S. Pakhomov, and V. A. Shchurov, J.\nExp. Theor. Phys. 29, 669 (1969).\n[45] P. P. Maksimenkov and V. I. Ozhogin, J. Exp. Theor.\nPhys. 38, 324 (1974).\n[46] F. Crameri, Zenodo 10.5281/ZENODO.5501399.\n[47] F. Crameri, G. E. Shephard, and P. J. Heron, Nat. Com-\nmun. 11, 5444 (2020)." }, { "title": "1112.1646v1.Magnetization_and_spin_gap_in_two_dimensional_organic_ferrimagnet_BIPNNBNO.pdf", "content": "arXiv:1112.1646v1 [cond-mat.mes-hall] 7 Dec 2011Magnetization and spin gap in two-dimensional organic\nferrimagnet BIPNNBNO\nV.E. Sinitsyn, I.G. Bostrem, A.S. Ovchinnikov\nDepartment of Physics, Ural State University, 620083, Ekat erinburg, Russia\nY. Hosokoshi\nDepartment of Physical Science, Osaka Prefecture Universi ty, Osaka, Japan\nK. Inoue\nDepartment of Chemistry, Hiroshima University, Hiroshima, Japan\n(Dated: November 9, 2018)\nA magnetization process in two-dimensional ferrimagnet BI PNNBNO is analyzed.\nThe compound consists of ferrimagnetic (1,1/2) chains coup led by two sorts of anti-\nferromagnetic interactions. Whereas a behavior of the magn etization curve in higher\nmagnetic fields can be understood within a process for the sep arate ferrimagnetic\nchain, an appearance of the singlet plateau at lower fields is an example of non-Lieb-\nMattis typeferrimagnetism. By usingthe exact diagonaliza tion technique for a finite\nclusters of sizes 4 ×8 and 4×10 we show that the interchain frustration coupling\nplays an essential role in stabilization of the singlet phas e. These results are comple-\nmented by an analysis of four cylindrically coupled ferrima gnetic (1,1/2) chains via\nan abelian bosonization technique and an effective theory bas ed on the XXZ spin-\n1/2 Heisenberg model when the interchain interactions are s ufficiently weak/strong,\nrespectively.\nI. INTRODUCTION\nDuring the last fifteen years, two-dimensional (2D) quantum spin s ystems have attracted\na lot of attention both from theoretical and experimental physicis ts. A competition between\nconventional classically ordered phases and more exotic quantum o rdered phases lies in\nthe focus of the investigations. Magnetic systems with a finite corr elation length at zero\ntemperatureandafinitespingapabovethesingletgroundstate, s pinliquids, realizeHaldane2\nprediction at the level of two space dimensions1.\nTo date, one can distinguish two main routes in studies of 2D spin gap c ompounds. A\nformation of spin gap in spin dimer systems, for example SrCu 2(BO3)22and CaV 4O93, is\nexplained by a modified exchange topology similar to Shastry-Suther land lattice4. Another\nway to increase quantum fluctuations and stabilize a spin liquid ground state is realized\nin kagome antiferromagnet5. Experimental candidates for 2D kagome antiferromagnets are\ncurrently available: herbertsmithite6–8and volborthite9,10. Both these strategies deal with\nantiferromagnetic compounds. In view of this, an observation of a singlet ground state with\na pronounced spin gap in 2D ferrimagnetic material BIPNNBNO seems exotic12.\nThe crystal structure of BIPNNBNO is shown in Fig. 1. A magnetic un it of the spin\nsystem presents organic triradical BIPNNBNO. Each of the molecu les includes three s=1/2\na\nb\nFIG. 1: Magnetic model of BIPNNBNO crystal. The black (white ) circles denote spins S=1\n(s=1/2).\nspins (see Fig. 2) with intramolecular ferromagnetic JFand antiferromagnetic JAFinterac-\ntions. The magnitude of |JF| ∼300 K is very large, and two spins coupled ferromagnetically\nbehave as a S - 1 moiety. Ferrimagnetic chains are stretched along t he b-axis. There are two\nkinds of antiferromagnetic interchain interactions along the a-axis . One is between the s-1/2\nspins, which connects the nearest neighboring chains. The other is between S-1 species,\nwhich connects the next nearest neighboring chains and introduce s spin frustration.3\nS=1S=1/2 S=1/2\nJAFJAFJAF\nJFN+\nO- NO\nN N\nO OJAFJAF\nJF\nFIG. 2: Molecular structure of BIPNNBNO and the elementary m agnetic cell at JF≫ |JAF|.\nThe puzzle is following. It iswell known that a low-energy physics of an isolated ferrimag-\nnetic (S,s) chain corresponds to a gapless (S-s) ferromagnet13. Quite predictably one might\nexpect an appearance of an ordered state with fluctuations in the form of spin waves near\nthe classical state. However, measurements of magnetization sh ows that an opening of a gap\nby analogy with the Haldane chain is likely scenario. Such a behavior is a m anifestation of\nnon-Lieb-Mattis type ferrimagnetism11. Namely, the magnetization measured at 400 mK is\nnearly zero below 4.5 T, increases rapidly above 4.5 T and exhibits a bro ad 1/3 plateau and\na narrow 2/3 one at 7-23 T and around 26 T, respectively. Above 29 T, the magnetization\nis completely saturated12.\nThepurposeofthepaperistoinvestigatethemagnetizationproce ss. Theproblemiscom-\nplicated by a lack of reliable information about intra- and interchain ex change interactions.\nSo, before studying of the 2D ferrimagnetic system BIPNNBNO we d evelop a simple quan-\ntum mechanical approach that models a magnetization process of t he ferrimagnetic chain\n(1,1/2). The treatment agrees qualitatively with a predictions of th e theory for quantum\nspin chains14and provides reasonable estimations of the exchange intrachain couplings. In\naddition, it captures a peculiarity of a magnetization process in the p rototype 2D material,\ni.e. an appearance of the intermediate 2/3 plateau. Given these est imations we examine the\nmagnetization process in BIPNNBNO by analyzing exact diagonalizatio n (ED) calculations\nfor a finite clusters of size N= 32 andN= 40. A main conclusion to be drawn from these\ncalculations that an emergence of the anomalous singlet plateau is a c onsequence of the\nfrustrating interchain interaction.\nTwo different mechanisms of formation of the plateau may be likely can didates: a gener-4\nalization of Haldanes conjecture to the weakly coupled ferrimagnet ic chains, and a valence-\nbond (dimerized) type ground state in the strong-coupling limit. To d etermine what of these\nscenarios is relevant we develop low-energy effective theories for t he 4-legs spin tube, which\nforms a minimal setup including the interchain couplings. In the regime of weakly coupled\nspin tube legs we apply abelian bosonization technique. The opposite lim it of a strong-ring\ninteraction is analyzed in terms of an effective Heisenberg XXZ model where the intrachain\ncoupling is perturbatively taken into account. Our analytical treat ment shows that only the\nfirst approach confirms an important role of frustration in stabiliza tion of the singlet phase.\nNotethatastudyofspintubesisofinterest byitselfbecauseboth offrustrationandquan-\ntum fluctuation arestrong15. Our model isdirectly related with thecompound BIPNNBNO,\nbut the main results are expected to apply to other frustrated sp in tubes as well. Re-\ncently, it has been reported that the experimental candidate for the four-leg spin tube,\nCu2Cl4·D8C4SO2, is available16.\nThe paper is organized as follows: In Sec. II we consider a magnetiza tion process in the\nferrimagnetic chain (1 ,1/2). In Sec. III we discuss results of magnetization process in two-\ndimensional ferrimagnetic system obtained via the exact diagonaliza tion method on a finite\ncluster. In Sec. IV we derive effective low-energy spin-1/2 Hamilton ian. A bosonisation\nstudy of the spin tube is carried out in Sec. V. A discussion of these r esults is relegated to\nthe Conclusion part.\nII. MAGNETIZATION OF AN ISOLATED FERRIMAGNETIC CHAIN\nThe issue that we address below is whether a calculation for an isolate d ferrimagnetic\nchain partially reproduces features of the magnetization curve ob served in the BIPNNBNO\ncrystal. We demonstrate that both the 1/3 plateau and the 2/3 pla teau can be recovered\nwithin a simple quantum mechanical analysis of a magnetization proces s of an isolated\nquantum (1 ,1/2) ferrimagnetic chain under an applied magnetic field. This enables to\nestimate the intrachain exchange parameters JAFandJ1.5\nS=N/2\nS=N/2+1\nS=N\nS=N-1\nS=N+1\nS=3N/2-1\nS=3N/2\nFIG. 3: Schematic picture of states of the ferrimagnetic cha in used in construction of a magneti-\nzation curve. Excited blocks are marked by the gray shadow.\nWe start from the values of the critical fields derived from the ED da ta17\n\n\nB1=∂ε\n∂m/vextendsingle/vextendsingle\nm→1/3+0≈E(N/2+1)−E(N/2),\nB2=∂ε\n∂m/vextendsingle/vextendsingle\nm→2/3−0≈E(N)−E(N−1),\nB3=∂ε\n∂m/vextendsingle/vextendsingle\nm→2/3+0≈E(N+1)−E(N),\nBsat=∂ε\n∂m/vextendsingle/vextendsingle\nm→1−0≈E(3N/2)−E(3N/2−1),(1)\nwhereε,mare the energy and the magnetization per an elementary cell of the (1,1/2)\nferrimagnetic chain, Nis a number of the elementary cells.\nTo estimate the energies E(S) in the right-hand side of Eqs.(1), where Sis a total spin\nof the chain, we use the Hamiltonian of the one-dimensional quantum (1,1/2) ferrimagnet\nˆHc=JAFN/summationdisplay\ni=1/vectorS1i/vector s2i+J1N/summationdisplay\ni=1/vector s2i/vectorS1i+1(S1= 1, s2= 1/2), (2)6\nand construct the required quantum states |SM/an}bracketri}htwith the given quantum numbers of the\ntotal spinSand itsz-projection M.\nWe suppose that the 1 /3 magnetization plateau corresponds to the ground state of the\n(1,1/2) ferrimagnetic chain with S=N/2. The wave function of the polarized state is given\nby\n|N/2,N/2/an}bracketri}ht=N/productdisplay\ni=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\ni, (3)\nand presents a direct product of the spin states of the magnetic e lementary cells (1,1/2) [see\nFig. 3]. The corresponding energy eigenvalue equals to\nE(N/2) =−JAFN−1\n9J1N. (4)\nWith an increasing of a magnetic field the ground state (3) is destroy ing and the state\nwithS=N/2 + 1 stabilizes. A low-lying excitation may be qualitatively considered as a\nforming of one triplet bond. The trial new wave function is\n|N/2+1,N/2+1/an}bracketri}ht=1√\nNN/summationdisplay\nk=1\nN/productdisplay\ni(/negationslash=k)=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\ni\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\nk=N/summationdisplay\nk=1αkΨk.(5)\nIt is composed from all arrangements of the excited block within the chain taken with equal\nweightsαk.\nBy introducing the state and calculating the matrix element\n/an}bracketle{tΨk|ˆHc|Ψk′/an}bracketri}ht=/bracketleftbigg\nE(N/2)+3\n2JAF+7\n18J1/bracketrightbigg\nδkk′−1\n3J1δk,k′±1 (6)\none obtains the relationship for the coefficients αk\n−1\n3J1αk−1+/bracketleftbigg\nE(N/2)+3\n2JAF+7\n18J1−E(N/2+1)/bracketrightbigg\nαk−1\n3J1αk+1= 0,(7)\nwhich is tantamount to\nE(N/2+1) =E(N/2)+3\n2JAF+7\n18J1−1\n3J1/parenleftbiggαk−1\nαk+αk+1\nαk/parenrightbigg\n. (8)\nThis expression includes two independent variational parameters αk−1/αkandαk+1/αk. The\nminimal value\nEmin(N/2+1) =E(N/2)+3\n2JAF−5\n18J1 (9)7\nis reached provided αk−1/αk=αk+1/αk= 1. This yields the critical magnetic field B1\ndestroying the 1 /3 plateau\nB1=3\n2JAF−5\n18J1. (10)\nTo find the critical fields B2andB3of the beginning and the end of the 2/3 plateau,\nrespectively, we construct the trial states\n|N,N/an}bracketri}ht=N/2/productdisplay\ni=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2i−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2i, (11)\n|N−1,N−1/an}bracketri}ht=1√\nNN/2/summationdisplay\nk=1/bracketleftBiggk−1/productdisplay\ni=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2i−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2i/bracketrightBigg\n×/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2k−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2k\nN/2/productdisplay\ni=k+1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2i−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2i\n,(12)\n|N+1,N+1/an}bracketri}ht=1√\nNN/2/summationdisplay\nk=1/bracketleftBiggk−1/productdisplay\ni=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2i−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2i/bracketrightBigg\n×/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2k−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2k\nN/2/productdisplay\ni=k+1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\n2i−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\n2i\n,(13)\nwhich are schematically shown in Fig. 3.\nBy the same manner we obtain\nB2=3\n2JAF+7\n18J1, B3=3\n2JAF+5\n6J1. (14)\nThe saturation field Bsatis determined with an aid of the trial wave functions\n|3N/2,3N/2/an}bracketri}ht=N/productdisplay\ni=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\ni, (15)\n|3N/2−1,3N/2−1/an}bracketri}ht=1√\nNN/summationdisplay\nk=1\nN/productdisplay\ni(/negationslash=k)=1/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg3\n23\n2/angbracketrightbigg\ni\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle/parenleftbigg\n11\n2/parenrightbigg1\n21\n2/angbracketrightbigg\nk(16)\nthat results in\nBsat=3\n2JAF+3\n2J1. (17)\nGiven the experimental estimations for the 2D BIPNNBNO system, B1∼31K,B2≈\nB3∼35K, andBsat= 39K, we obtain from Eqs.(10,17) the values of the intrachain\nexchange couplings, JAF≈21KandJ1≈3.5K. By substituting them into Eq.(14) we get8\nthe critical fields of the 2/3 plateau, B2≈32.8K(24.4T) andB3≈34.4K(25.6T). A\nqualitativebehavioroftheferrimagneticchainmagnetizationcurve builtfromthesereference\npoints is depicted in Fig. 4. We emphasize especially that an emergence of the intermediate\n2/3 plateau is not related with interchain frustration effects.\n/s48 /s53 /s49/s48 /s49/s53 /s50/s48 /s50/s53 /s51/s48 /s51/s53 /s52/s48/s48/s44/s50/s48/s44/s52/s48/s44/s54/s48/s44/s56/s49/s44/s48\n/s49/s47/s51/s50/s47/s51/s77/s47/s77\n/s115/s97/s116\n/s66/s40/s84/s41\nFIG. 4: Qualitative magnetization curve of the ferrimagnet ic (1,1/2) chain.\nIII. MAGNETIZATION: EXACT DIAGONALIZATION\nnnJ\nnnnJAFJ 1J\nFIG. 5: Cluster of N= 32 sites used in the exact diagonalization. The fixed exchan ge couplings\nareJAF= 21K,J1= 3.5K,Jnn= 0.5J1= 1.75K.\nIn order to understand a role of the interchain couplings the magne tization process of\nthe BIPNNBNO ferrimagnet was examined by the variant of a numeric al diagonalization\nmethod with conservation of a total cluster spin18,19.9\nThe model Hamiltonian is given by\nˆHclust=JAF/summationdisplay\nij/vectorSi/vector sj+J1/summationdisplay\nij/vectorSi/vector sj+Jnn/summationdisplay\nij/vector si/vector sj+Jnnn/summationdisplay\nij/vectorSi/vectorSj, (18)\nwhere/vectorSi(/vector si) denotes spin-1 (spin-1/2) operator at site i. The sublattices and the network\nof the antiferromagnetic interactions, JAF,J1,JnnandJnnn, are shown in Fig. 5. We\nperform calculation of the N-step magnetization curve for the N=3 2 cluster depicted in the\nsame Figure. The intrachain parameters, JAFandJ1, have been estimated in the previous\nSection whereas the interchain ones, JnnandJnnn, are assumed to be less than J1. The open\nboundary conditions are used for the numerical calculations.\nThe magnetization process is compared with the results of the mode l of non-interacting\n(1,1/2) ferrimagnetic chains. A standard way to build magnetization curve atT= 0\nis to define the lowest energy E(N,M) of the Hamiltonian (2) in the subspace where\n/summationtextN\nj=1/parenleftbig\nSz\nj+sz\nj/parenrightbig\n=Mfor a finite system of Nelementary ( S,s) blocks. Applying a magnetic\nfieldBleads to a Zeeman splitting of the energy levels, and therefore level crossing occurs on\nincreasing the field. These level crossing correspond to jumps in th e magnetization until the\nfully polarized state is reached at a certain value of the magnetic field . The magnetization\nof four independent chains is then derived from\nm= 4M/N, M = max[M|E(N,M+1)−E(N,M)>B], (19)\nwhich gives a step curve.\nAn importance of the frustrating coupling is seen from comparison o f two magnetization\ncurves displayed in Figs. 6, 7. They correspond to no frustration c ase and a pronounced\nfrustrating coupling, respectively. The magnetization curves exh ibit several interesting fea-\ntures. For instance, the magnetization behavior in higher magnetic fields (B > B 1) is well\nreproduced within the model of non-interacting ferrimagnetic cha ins. Another remarkable\nfeature revealed by Figs. 6, 7 that the singlet ground state platea u emerges at non-zero\nfrustration interaction whereas the narrow 2 /3 plateau appears regardless of the frustration.\nWe numerically found that the width ∆ Sof the singlet plateau scales almost linearly with\naJnnnvalue (Fig. 8). To check into the case of the dependence we repeat cacluations on\na cluster of larger size, N= 40, with the same set of parameters that support the finding.\nThe observation points out that the zero magnetization plateau ha s a quantum origin with\na crucial role of frustration which destroys a long-range order an d drives the system into the10\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48/s48/s44/s48/s48/s44/s53/s49/s44/s48\n/s50/s47/s51\n/s49/s47/s51\n/s66/s40/s84/s41/s77/s47/s77\n/s115/s97/s116\nFIG. 6: Magnetization curve for the 32-site cluster. The exc hange couplings are taken as JAF=\n21K,J1= 3.5K,Jnn= 0.5J1, andJnnn= 0 (no frustration). The dotted line marks a calculation\nvia the model of non-interacting (1 ,1/2) chains.\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48/s48/s44/s48/s48/s44/s53/s49/s44/s48/s77/s47/s77\n/s115/s97/s116\n/s66/s40/s84/s41/s49/s47/s51/s50/s47/s51\nFIG. 7: Magnetization curve for the 32-site cluster. The exc hange couplings are taken as JAF=\n21K,J1= 3.5K,Jnn= 0.5J1,Jnnn= 0.075J1. The dotted line marks a calculation via the model\nof non-interacting (1 ,1/2) chains.11\n0□00. 0□02 . 0□04 . 0□06 . 0□08 . 0□10 . 0□12 . 0□14 .0□00.0□02.0□04.0□06.0□08. /c68S/JAF\nJnnn/JAF\nFIG. 8: Value of the singlet plateau ∆ Sas a function of the frustrating coupling Jnnnobtained on\na cluster of size N= 32 (black circles) and N= 40 (white circles).\nsinglet phase. Below, we address analytically the issue in the regimes o f strong and weak\ninterchain couplings.\nIV. A FORMATION OF THE SINGLET PLATEAU\nIn low-dimensional Heisenberg systems frustrating couplings can d rive transitions to gap-\nfull quantum states, where local singlets forma groundstate. Th ese quantum gappedphases\nmay have long-rangedsinglet order (valence bondstate), or realiz ea resonating valence band\nspin liquid. In last case, a ground state is a coherent superposition o f all lattice-coverings\nby local singlets21.\nTorecognizefeaturesofthesephasesintheEDresultsweundert akeanalytical treatments\nof the four-legs spin tube shown in Fig. 9. The new system is infinite alo ng theb-axis,\nand periodic with the 4-site period along the a-axis. The tube forms a minimal setup\nincluding the interchain nearest- and next-to-nearest neighbor c ouplings and contains the\nsame number of ferrimagnetic chains parallel to the b-axis as the clu sters in the ED study.\nAs we demonstrate below, the simplified model elucidates an importan t role of frustration\nin stabilization of the singlet phase.12\n1J\nnnJ\nnnnJ\nJ\nJJ’1\n2\n3412\n34\nFIG. 9: 4-leg spin tube structure used in the limit of strong r ing coupling (up) and in the case\nof weakly interacting chains (below). The black (white) cir cles mean spin-1 (spin-1/2) sites. The\ngray circles denote renormalized spin-1/2 blocks.\nThe singlet phase may arise in the limit of strong ring coupling Jnn,Jnnn≫J1. In this\ncase, the problem can be analyzed in terms of Heisenberg XXZ model similar to ladders in a\nmagnetic field20. The opposite limit ( J1≫Jnn,Jnnn) results in a scenario of weakly interact-\ning chains. Based on a block renormalization procedure the original s ystem is then mapped\nonto the model of a spin tube with four ferromagnetic spin-1/2 legs . We mention that the\nground state properties of two-leg spin ladders with ferromagnet ic intrachain coupling and\nantiferromagnetic interchain couplings have been discussed in Refs .30,31in an absence of an\nexternal field. A magnetization process of those spin ladders with a n even number of legs\n(2 and 4) has been studied in Ref.32in the regime of weak ferromagnetic coupling along the\nlegs and strong antiferromagnetic coupling along the rungs.\nAn appearance of the singlet phase in the frustrated spin tube with four weakly coupled\nferromagnetic spin-1/2 legs can be studied through the bosonizat ion technique which proves\neffectiveness for quasi-one-dimensional spin-one-half systems. To the best of our knowledge,\nthe system has never been previously reported, however our fur ther analysis follows closely\nto that of given in Ref.24, where 4-legs spin tube with antiferromagnetic chains and a specific13\nform of diagonal rung interactions (but with no frustration) has b een treated. Note as well\nthat spin ladders with ferromagnetic and ferrimagnetic legs are muc h less studied26,27by the\nbosonization approach in comparison with ladder systems with antife rromagnetic legs. The\nmain problem arising here is that the formalism is well defined only if ther e is an easy-plane\nexchange anisotropy. In this regard, we note that measurement s of the angular dependence\nof the ESR linewidth for the BIPNNBNO system showed that the large st linewidth was\nobserved for the field direction perpendicular to the ab plane28. Due to the theoretical\nconsideration by Oshikawa and Affleck29a critical regime of XY-anisotropy is expected in\nthe compound. In addition we point out that the ED algorithm invoked in the previous\nSection enables to treat clusters of sufficiently large sizes due to us e of the rotational SU(2)\nsymmetry. The latter is broken by the anisotropy whose role in a sing let gap formation is a\nsubject of future ED studies.\nA. Spin tube: weakly interacting rings and a model of a single XXZ chain\nWe study the Hamiltonian of the spin tube (see Fig. 9)\nH=N/summationdisplay\nn=1Hring\nn+J1N/summationdisplay\nn=1(Sn,1sn+1,1+sn,2Sn+1,2+Sn,3sn+1,3+sn,4Sn+1,4)−BN/summationdisplay\nn=14/summationdisplay\ni=1/parenleftbig\nSz\nn,i+sz\nn,i/parenrightbig\n,\n(20)\nwhere the Hamiltonian of the separate ring is\nHring\nn=Jnn(sn,1sn,2+sn,2sn,3+sn,3sn,4+sn,4sn,1)+JnnnN/summationdisplay\nn=1(Sn,1Sn,3+Sn,2Sn,4).\nHere,S= 1 ands= 1/2,nis the index of the ring, Nis the total number of rings, and the\nindeximarks the (1,1/2) blocks inside the rings. Periodic boundary conditio ns along the\ntube direction are imposed. In our model it is suggested that Jnn,Jnnn≫J1.\nIn the limit J1= 0 the system decouples into a collection of nonineracting rings. At z ero\nmagnetic field the singlet and triplet states\n|ψ0/an}bracketri}ht=−√\n3\n2|00;00/an}bracketri}ht+1\n2|11;00/an}bracketri}ht,\n|ψ1/an}bracketri}ht=1√\n2|01;11/an}bracketri}ht+1√\n2|10;11/an}bracketri}ht (21)\nhave the lowest energies E0=−2Jnn/9+8Jnnn/9 andE1=−Jnn/9+8Jnnn/9, respectively.\nThestatesofthering |S12S34;SM/an}bracketri}htareobtainedviathecommonruleofadditionofmoments,14\nwhereS12(S34) is spin of dimer composed of the spins of the 1 and 2 (3 and 4) blocks. The\nsinglet and triplet states of the ring that enter into (21) are given in Appendix A.\nUpon increasing the magnetic field a transition between the singlet an d triplet states\noccurs atB=Jnn/9 and the the total magnetization jumps abruptly from zero to M=N.\nAtnon-zeroringcouplingthesharptransitionisbroadenedandsta rtsfromacriticalvalue\nB0. To find the field we derive the XXZ spin chain Hamiltonian by we using the standard\napproach that is analogous to study of spin-1/2 ladder with strong rung exchange20.\nThe Hamiltonian (20) is splitted into two parts\nH=H0+H1.\nH0=N/summationdisplay\nn=1Hring\nn−Bc4/summationdisplay\ni=1N/summationdisplay\nn=1(Sz\nin+sz\nin),\nH1=J1N/summationdisplay\nn=1[Sn,1sn+1,1+sn,2Sn+1,2+Sn,3sn+1,3+sn,4Sn+1,4]−(B−Bc)4/summationdisplay\ni=1N/summationdisplay\nn=1(Sz\nin+sz\nin),\nwhereBc=E1−E0. TheH1lifts the 2N-fold degeneracy of the ground state of the\nHamiltonian H0. The later can be either in the state |ψ0/an}bracketri}htor|ψ1/an}bracketri}ht. By using the standard\nmany body perturbation theory22the effective Hamiltonian can be derived\nHeff=Jeff\nxyN/summationdisplay\nn=1/parenleftBig\n˜Sx\nn˜Sx\nn+1+˜Sy\nn˜Sy\nn+1/parenrightBig\n+Jeff\nzN/summationdisplay\nn=1˜Sz\nn˜Sz\nn+1−Heff\nzN/summationdisplay\nn=1˜Sz\nn, (22)\nwhereJeff\nxy=−16J1/27,Jeff\nz=−J1/9 andBeff\nz=J1/9+B−Bc.\nTo get the expression the pseudo-spin ˜Si= 1/2 operators that act on the states |ψ0/an}bracketri}htand\n|ψ1/an}bracketri}htare introduced\n˜Sz\nn|ψ0/an}bracketri}htn=−1\n2|ψ0/an}bracketri}htn,˜Sz\nn|ψ1/an}bracketri}htn=1\n2|ψ1/an}bracketri}htn,\n˜S+\nn|ψ0/an}bracketri}htn=|ψ1/an}bracketri}htn,˜S+\nn|ψ1/an}bracketri}htn= 0,\n˜S−\nn|ψ0/an}bracketri}htn= 0,˜S−\nn|ψ1/an}bracketri}htn=|ψ0/an}bracketri}htn. (23)\nThe starting spin-1 operators and the pseudo-spin operators in t he restricted space are\nrelated by\nSz\nin=1\n6+1\n3˜Sz\nn,\nS+\nin= (−1)i−14\n3√\n3˜S+\nn, S−\nin= (−1)i−14\n3√\n3˜S−\nn. (24)15\nThe corresponding map for the spin-1/2 operators is\nsz\nin=−1\n24−1\n12˜Sz\nn,\ns+\nin=(−1)i\n3√\n3˜S+\nn, s−\nin=(−1)i\n3√\n3˜S−\nn. (25)\nThe Jordan-Wigner transformation maps the Hamiltonian (22) onto a system of inter-\nacting spinless fermions\nHsf=tN/summationdisplay\nn=1/parenleftbig\nc+\nici+1+c+\ni+1ci/parenrightbig\n+VN/summationdisplay\nn=1nini+1−µN/summationdisplay\nn=1ni, (26)\nwheret=Jeff\nxy/2,V=Jeff\nzandµ=Jeff\nz+Beff\nz.\nThe lowest critical field B0corresponds to that value of thechemical potential µfor which\nthe band of spinless fermions starts to fill up. This yields the conditio nµ=−2tand leads\nto the result\nB0=1\n9Jnn+16\n27J1. (27)\nThe critical value involves no frustration parameter Jnnnthat is clearly contrary to the ED\nresults.\nA similar analysis can be carried out for the saturation field. The deta ils of the calcula-\ntions are relegated to Appendix B.\nB. Spin tube: a model of weakly interacting ferromagnetic legs and abelian\nbosonization\nTo apply the bosonization we should map the initial system, consisting of two sorts of\nspins, spin-1/2 and spin-1, to the spin-1/2 system by using the qua ntum renormalization\ngroup (QRG) in real space based on the block renormalization proce dure23. To exploit\nthe real-space QRG technique one divide the spin lattice into small bloc ks, namely, the\nintrachain dimers (1,1/2), and obtains the lowest energy states {|α/an}bracketri}ht}of each isolated block.\nThe effect of inter-block interactions is then taken into account by constructing an effective\nHamiltonian Heffwhich now acts on a smaller Hilbert space embedded in the original one.\nIn this new Hilbert space each of the former blocks is treated as a sin gle site. The effective\nHamiltonians Heff=Q†HQis constructed via the projection operator Q=N/producttext\ni=1Qiwith16\nQi=m/summationtext\nα=1|α/an}bracketri}ht/an}bracketle{tα|of eachi-th block where mis the number of low energy states that are kept\nandNis a number of lattice cells.\nWe hold the lowest doublet S−1/2 to find an effective low-energy Hamiltonian. The\nhigher energy S−3/2 states are neglected. One can check that the reduced matrix ele ments\nare (S1= 1,s2= 1/2)\n/an}bracketle{t11/2;1/2||S1||11/2;1/2/an}bracketri}ht= 2/radicalbigg\n2\n3,\n/an}bracketle{t11/2;1/2||s2||11/2;1/2>=−1√\n6.\nTherefore the effective spin-1/2 operators of the renormalized c hain are\nQ†\ni/vectorS1iQi=4\n3/vectorSi, Q†\ni/vector s2iQi=−1\n3/vectorSi(S= 1/2). (28)\nThe renormalized Hamiltonian of the intrachain interactions corresp onds to the ferromag-\nnetic Heisenberg spin-1/2 model with the exchange coupling J=−4J1/9 (Fig. 9). The\ninterchain interactions between the nearest neighbors, spins -1/ 2, and next to the nearest\nneighbors, spins -1, are renormalized as J⊥=Jnn/9 andJ′\n⊥= 16Jnnn/9, respectively.\nConsider a four-legs spin tube consisting of spin-1 /2 chains. The Hamiltonian of the\nsystem is\nˆHtube=4/summationdisplay\nλ=1ˆHλ+ˆH⊥\n12+ˆH⊥\n23+ˆH⊥\n34+ˆH⊥\n14+ˆH′⊥\n13+ˆH′⊥\n24. (29)\nThe spins along the chains are coupled ferromagnetically, the Hamilto nian for the separate\nλ-th chain is\nˆHλ=−JxyN/summationdisplay\ni=1/parenleftbig\nSx\nλ,jSx\nλ,j+1+Sy\nλ,jSy\nλ,j+1/parenrightbig\n−JzN/summationdisplay\ni=1Sz\nλ,jSz\nλ,j+1,\nwhereSx,y,z\nλ,jare the spin S=1/2 operators at the jth site, the intraleg coupling is ferromag-\nnetic,J >0.\nThe interaction parts are given by\nˆH⊥\nλλ′=Jxy\n⊥,λλ′N/summationdisplay\nj=1/parenleftbig\nSx\nλ,jSx\nλ′,j+Sy\nλ,jSy\nλ′,j/parenrightbig\n+Jz\n⊥,λλ′N/summationdisplay\ni=1Sz\nλ,jSz\nλ,j, (30)\nˆH′⊥\nλλ′=J′xy\n⊥,λλ′N/summationdisplay\nj=1/parenleftbig\nSx\nλ,jSx\nλ′,j+Sy\nλ,jSy\nλ′,j/parenrightbig\n+J′z\n⊥,λλ′N/summationdisplay\ni=1Sz\nλ,jSz\nλ,j (31)17\nandincludes thenearest, J⊥>0, andthenext-to-nearest, J′\n⊥>0, antiferromagneticinterleg\ncouplings.\nThe unitary transformation keeping spin commutation relations\nSx,y\nλ,j→(−1)jSx,y\nλ,j, Sz\nλ,j→Sz\nλ,j\nmaps the Hamiltonian (29) to the Hamiltonian with antiferromagnetic le gs. It changes\nJxy→ −JxyandJz→Jz, and the ferromagnetic isotropic point is ∆ = Jz/Jxy=−1 in\nthe Hamiltonian with the antiferromagnetic legs.\nFollowing the general procedure of transforming a spin model to an effective model of\ncontinuum field, we convert the spin Hamiltonian of the spin tube with antiferromagnetic\nlegs to a Hamiltonian of spinless fermions using Jordan-Wigner transf ormation, then map\nit to a modified Luttinger model. The bosonic expressions for spin ope rators are\nS+\nλ(x) =S+\njλ\na=e−i√πΘλ\n√\n2πa/bracketleftBig\ne−i(πx/a)+cos/parenleftBig√\n4πΦλ/parenrightBig/bracketrightBig\n,\nSz\nλ(x) =Sz\njλ\na=1√π∂xΦλ+1\nπaei(πx/a)sin/parenleftBig√\n4πΦλ/parenrightBig\n, (32)\nwhere Φ and Θ are the bosonic dual fields, and xis defined on the lattice, xj=ja,ais a\nshort-distance cutoff.\nThe bosonized form of the Hamiltonian of the non-interacting chains is\nHλ=u\n2/integraldisplay\ndx/bracketleftbigg\nKΠ2\nλ+1\nK(∂xΦλ)2/bracketrightbigg\n, (33)\nwhere Π λ(x) =∂xΘλis canonically conjugate momentum to Φ λ. The Luttinger liquid\nparameters are fixed from the Bethe ansatz solution33\nK=π\n2(π−arccos∆), u=Jxyπ√\n1−∆2\n2arccos∆. (34)\nThe velocity uvanishes and Kdiverges for ∆ = −1. This corresponds to the ferromagnetic\ninstability point of a single chain.\nThe interchain interactions (30) between the nearest neighbor ch ains reads as\nH⊥\nλλ′=Jz\n⊥,λλ′/integraldisplaydx\nπ(∂xΦλ)(∂xΦλ′)+g1/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4π(Φλ+Φλ′)/parenrightBig\n+g2/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4π(Φλ−Φλ′)/parenrightBig\n+g3/integraldisplaydx\n(2πa)2cos/parenleftbig√π(Θλ−Θλ′)/parenrightbig18\n+g4/integraldisplaydx\n(2πa)2cos/parenleftbig√π(Θλ−Θλ′)/parenrightbig\ncos/parenleftBig√\n4π(Φλ+Φλ′)/parenrightBig\n+g5/integraldisplaydx\n(2πa)2cos/parenleftbig√π(Θλ−Θλ′)/parenrightbig\ncos/parenleftBig√\n4π(Φλ−Φλ′)/parenrightBig\n, (35)\nwhereg1=−2Jz\n⊥,λλ′,g2= 2Jz\n⊥,λλ′,g3= 2πJxy\n⊥,λλ′,g4=g5=πJxy\n⊥,λλ′. The Hamiltonian (31)\nof the next-to-nearest couplings has a similar form with a formal ch angeJz\n⊥,λλ′→J′z\n⊥,λλ′,\ng1→g′\n1=−2J′z\n⊥,λλ′,g2→g′\n2= 2J′z\n⊥,λλ′etc.\nFollowing the route of Refs.24,25it is convenient to introduce a symmetric mode Φ sand\nthree antisymmetric ones Φ a1, Φa2, Φa3\nΦs=1\n2(Φ1+Φ2+Φ3+Φ4),\nΦa1=1\n2(Φ1+Φ2−Φ3−Φ4),\nΦa2=1\n2(Φ1−Φ2−Φ3+Φ4),\nΦa3=1\n2(Φ1−Φ2+Φ3−Φ4).(36)\nIn terms of the new fields the quadratic part of the Hamiltonian (29) is diagonalized to\nH0=us\n2/integraldisplay\ndx/bracketleftbigg\nKsΠ2\ns+1\nKs(∂xΦs)2/bracketrightbigg\n+3/summationdisplay\ni=1uai\n2/integraldisplay\ndx/bracketleftbigg\nKaiΠ2\nai+1\nKai(∂xΦai)2/bracketrightbigg\n(37)\nwith\nus=u/parenleftbigg\n1+2KJz\n⊥\nuπ+KJ′z\n⊥\nuπ/parenrightbigg1\n2\n, K s=K/parenleftbigg\n1+2KJz\n⊥\nuπ+KJ′z\n⊥\nuπ/parenrightbigg−1\n2\n,\nua1=ua2=u/parenleftbigg\n1−KJ′z\n⊥\nuπ/parenrightbigg1\n2\n, K a1=Ka2=K/parenleftbigg\n1−KJ′z\n⊥\nuπ/parenrightbigg−1\n2\n,\nua3=u/parenleftbigg\n1−2KJz\n⊥\nuπ+KJ′z\n⊥\nuπ/parenrightbigg1\n2\n, K a3=K/parenleftbigg\n1−2KJz\n⊥\nuπ+KJ′z\n⊥\nuπ/parenrightbigg−1\n2\n.(38)\nThe relevant and marginally relevant terms of the interchain coupling s are given by\nHint= 2g12/summationdisplay\ni=1/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4πΦs/parenrightBig\ncos/parenleftBig√\n4πΦai/parenrightBig\n+2g′\n1/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4πΦs/parenrightBig\ncos/parenleftBig√\n4πΦa3/parenrightBig\n+2g22/summationdisplay\ni=1/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4πΦai/parenrightBig\ncos/parenleftBig√\n4πΦa3/parenrightBig\n+2g′\n2/integraldisplaydx\n(2πa)2cos/parenleftBig√\n4πΦa1/parenrightBig\ncos/parenleftBig√\n4πΦa2/parenrightBig\n+2g32/summationdisplay\ni=1/integraldisplaydx\n(2πa)2cos/parenleftbig√πΘai/parenrightbig\ncos/parenleftbig√πΘa3/parenrightbig\n+2g′\n3/integraldisplaydx\n(2πa)2cos/parenleftbig√πΘa1/parenrightbig\ncos/parenleftbig√πΘa2/parenrightbig\n+g4/integraldisplaydx\n(2πa)2/bracketleftBig\ncos/parenleftbig√π(Θa2+Θa3)/parenrightbig\ncos/parenleftBig√\n4π(Φs+Φa1)/parenrightBig19\n+cos/parenleftbig√π(Θa1−Θa3)/parenrightbig\ncos/parenleftBig√\n4π(Φs−Φa2)/parenrightBig\n+cos/parenleftbig√π(Θa2−Θa3)/parenrightbig\ncos/parenleftBig√\n4π(Φs−Φa1)/parenrightBig\n+cos/parenleftbig√π(Θa1+Θa3)/parenrightbig\ncos/parenleftBig√\n4π(Φs+Φa2)/parenrightBig/bracketrightBig\n+g′\n4/integraldisplaydx\n(2πa)2/bracketleftBig\ncos/parenleftbig√π(Θa1+Θa2)/parenrightbig\ncos/parenleftBig√\n4π(Φs+Φa3)/parenrightBig\n+cos/parenleftbig√π(Θa1−Θa2)/parenrightbig\ncos/parenleftBig√\n4π(Φs−Φa3)/parenrightBig/bracketrightBig\n. (39)\nTheg5terms are irrelevant and are omitted.\nThe Hamiltonian (37) describes four independent gapless spin-1 /2 chains coupled by the\ninterchain interaction in the form of Eq.(39). It is expected that th e interleg coupling results\nin the Haldane gap in the excitation spectrum. Note that in the vicinity of the single chain\nferromagnetic instability, ∆ = −1, the effective bandwidth collapses, u→0, and the effect\nof the interleg couplings becomes crucial. To find a detail behavior of the gap in the phase\ndiagram with antiferromagnetic ( J⊥,J′\n⊥>0) interleg coupling and ferromagnetic leg regime\n(∆<0) we use the renormalization group analysis.\nThe RG equations are derived through the standard technique (se e Ref.34, for example).\nThe result is\ndg1\ndl= [2−(Ks+Ka1)]g1,\ndg′\n1\ndl= [2−(Ks+Ka3)]g′\n1,\ndg2\ndl= [2−(Ka3+Ka1)]g2,\ndg′\n2\ndl= [2−2Ka1]g′\n2,\ndg3\ndl=/bracketleftbigg\n2−1\n4/parenleftbigg1\nKa1+1\nKa3/parenrightbigg/bracketrightbigg\ng3,\ndg′\n3\ndl=/bracketleftbigg\n2−1\n2Ka1/bracketrightbigg\ng′\n3,\ndg4\ndl=/bracketleftbigg\n2−/parenleftbigg\nKs+Ka1+1\n4Ka1+1\n4Ka3/parenrightbigg/bracketrightbigg\ng4,\ndg′\n4\ndl=/bracketleftbigg\n2−/parenleftbigg\nKs+Ka3+1\n2Ka1/parenrightbigg/bracketrightbigg\ng′\n4,\ndKs\ndl=−4g2\n1/parenleftbiggKs\n4πus/parenrightbigg2\n−2g′\n12/parenleftbiggKs\n4πus/parenrightbigg2\n−2g2\n4/parenleftbiggKs\n4πus/parenrightbigg2\n−g′\n42/parenleftbiggKs\n4πus/parenrightbigg2\n,\ndKa1\ndl=−1\n2g2\n1/parenleftbiggKa1\n2πua1/parenrightbigg2\n−1\n2g2\n2/parenleftbiggKa1\n2πua1/parenrightbigg2\n−1\n2g′\n22/parenleftbiggKa1\n2πua1/parenrightbigg2\n−1\n4g2\n4/parenleftbiggKa1\n2πua1/parenrightbigg220\n+1\n2/parenleftbiggg3\n4πua1/parenrightbigg2\n+1\n2/parenleftbiggg′\n3\n4πua1/parenrightbigg2\n+1\n4/parenleftbiggg4\n4πua1/parenrightbigg2\n+1\n4/parenleftbiggg′\n4\n4πua1/parenrightbigg2\n,\ndKa3\ndl=−1\n2g′\n12/parenleftbiggKa3\n2πua3/parenrightbigg2\n−g2\n2/parenleftbiggKa3\n2πua3/parenrightbigg2\n−1\n4g′\n42/parenleftbiggKa3\n2πua3/parenrightbigg2\n+/parenleftbiggg3\n4πua3/parenrightbigg2\n+1\n2/parenleftbiggg4\n4πua3/parenrightbigg2\n. (40)\nOne sees that the g1terms are relevant for Ks+Ka1<2; theg′\n1term is relevant for\nKs+Ka3<2; theg2terms are relevant for Ka3+Ka1<2; theg′\n2term is relevant for\nKa1<1; theg3term is relevant for K−1\na1+K−1\na3<8; theg′\n3term is relevant for Ka1>1/4.\nDespite the g4andg′\n4terms are irrelevant they are the most relevant terms which couple\nthe symmetric and antisymmetric modes24.\nUsing the RG equations the behavior of the gap in the whole phase diag ram can be\nestablished. Following to standard routine, we analyze the effect of the transversal ( Jxy\n⊥)\nand the longitudinal ( Jz\n⊥) parts of the interleg coupling separately.\n1. Transversal part of the interleg interactions\nIn this case Jxy\n⊥,J′xy\n⊥/ne}ationslash= 0 andJz\n⊥,J′z\n⊥= 0, the initial values of the coupling constants are\ngiven byg1(l= 0) =g′\n1(l= 0) = 0,g2(l= 0) =g′\n2(l= 0) = 0,g3(l= 0) = 2πJxy\n⊥,g′\n3(l= 0) =\n2πJ′xy\n⊥,g4(l= 0) =πJxy\n⊥andg′\n4(l= 0) =πJ′xy\n⊥. The bare Luttinger parameters are us=u,\nua1=ua2=ua3=u, andKs(l= 0) =K,Ka1(l= 0) =Ka2(l= 0) =Ka3(l= 0) =K.\nThe termg3is relevant for −1≤∆≤0 while the g4term is irrelevant. It is easily checked\nnumerically that g3,g′\n3grow whereas g4,g′\n4decrease under the RG. It means that Θ 1, Θ2, Θ3\nare locked in one of the vacuum states (Θ 1= Θ2= 0, Θ 3=√πor Θ1= Θ2=√π, Θ3= 0\nprovidedJxy\n⊥>J′xy\n⊥), fluctuations of the fields Θ ai(i= 1,2,3) are completely suppressed.\nTherefore arbitrary Jxy\n⊥> J′xy\n⊥/ne}ationslash= 0 generate a gap in the antisymmetric modes (Θ iare\npinned, disordered).\nAfter the fluctuations of the fields Θ iare stopped, the infrared behavior of the symmetric\nmode is governed by the term of the coupling with the antisymmetric m odes\n˜Hint= 2˜g42/summationdisplay\ni=1/integraldisplaydx\n(2πa)2cos(√\n4πΦs)cos(√\n4πΦai)\n+2˜g′\n4/integraldisplaydx\n(2πa)2cos(√\n4πΦs)cos(√\n4πΦa3). (41)21\nwhere\n˜g4= ¯g4/an}bracketle{tcos/parenleftbig√π[Θa1+Θa3]/parenrightbig\n/an}bracketri}ht= ¯g4/an}bracketle{tcos/parenleftbig√π[Θa2+Θa3]/parenrightbig\n/an}bracketri}ht,\n˜g′\n4= ¯g′\n4/an}bracketle{tcos/parenleftbig√π(Θa1+Θa2)/parenrightbig\n/an}bracketri}ht,\nand ¯g4, ¯g′\n4are renormalized couplings provided g3,g′\n3=O(1). Here, the invariance of ˜Hint\ngiven by Eq.(39) under Θ ai→ −Θaiyields/an}bracketle{tcos/parenleftbig√π(Θai−Θaj)/parenrightbig\n/an}bracketri}ht=/an}bracketle{tcos/parenleftbig√π(Θai+Θaj)/parenrightbig\n/an}bracketri}ht.\nDespiteeiΦaihas exponentially decaying correlations due to the Θ aiare pinned, a scrupulous\nanalysis24shows that the effective Hamiltonian for Φ spresents a standard sine-Gordon\nHamiltonian\n˜Heff=us\n2/integraldisplay\ndx/bracketleftbigg\n¯KsΠ2\ns+1\n¯Ks(∂xΦs)2/bracketrightbigg\n+g/integraldisplaydx\n(2πa)2cos/parenleftBig√\n16πΦs/parenrightBig\n,(42)\nwhere¯Ksis a renormalized value of K, andgis a new effective coupling constant. From\nthe correlation function /an}bracketle{texp/bracketleftbig\ni√\n16πΦs(x)/bracketrightbig\nexp/bracketleftbig\ni√\n16πΦs(y)/bracketrightbig\n/an}bracketri}ht= (a2/|x−y|2)4¯Ksit follows\nthat thegterm has a scale dimension 4 ¯Ks. Therefore, it is relevant for ¯Ks<1/2, when Φ s\nis pinned, i.e. becomes massive35.\nTo summarize, the transversal part of the interleg coupling suppo rts gapped antisymmet-\nric modes, the symmetric sector is gapped at ¯Ks<1/2, and remains gapless at ¯Ks>1/2.\nThe condition ¯Ks= 1/2 determines a boundary between the gapless Spin Liquid XY1\nphase36, and a generalization of the gapped Rung Singlets phase37for the for-leg spin tube\n(Fig.10). (Hereinafter, we retain names of phases used in the theo ry of spin ladders with\nferromagnetic legs.) In thelast case, spins onthe same rungs, or a longthe shortest diagonals\nform singlet pairs by a dynamical way.\n2. Longitudinal part of the interleg interactions\nFor the case of the longitudinal part of the interleg exchange, Jxy\n⊥,J′xy\n⊥= 0 andJz\n⊥,J′z\n⊥/ne}ationslash=\n0, the bare values of the coupling constants are given by g1(l= 0) =−2Jz\n⊥,g′\n1(l= 0) =\n−2J′z\n⊥,g2(l= 0) = 2Jz\n⊥,g′\n2(l= 0) = 2J′z\n⊥,g3(l= 0) =g′\n3(l= 0) = 0, and g4(l= 0) =g′\n4(l=\n0) = 0.\nThe strong-coupling phase diagramin the vicinity of ferromagnetic in stability point (∆ =\n−1andJz\n⊥,J′z\n⊥= 0)obtainedintheRGanalysisisshowninFig. 11. Inthesectordenot edas\nspin liquid II phase37theg1,2,g′\n1,2terms are irrelevant. The symmetric and antisymmetric\nmodes remain gapless. In the sector marked as a Haldane phase the termsg1,2,g′\n1,2are22\n/s48/s44/s48 /s48/s44/s53 /s49/s44/s48/s45/s49/s44/s48/s48/s45/s48/s44/s57/s56\n/s82/s117/s110/s103/s32/s83/s105/s110/s103/s108/s101/s116/s115\n/s74 /s47/s74/s32/s74 /s39/s47/s74 /s61/s48/s46/s50\n/s83/s112/s105/s110/s32/s76/s105/s113/s117/s105/s100/s32/s88/s89/s49/s32/s80/s104/s97/s115/s101\nFIG. 10: The ground-state phase diagram in the vicinity ∆ = −1 of the four-leg tube with\ntransverse coupling between legs.\nrelevant. Since all of the modes are coupled and locked together, b oth the symmetric and\nantisymmetric modes are gapped.\nThe phase of a ferromagnet with antiphase interchain order arises as a result of the\nferromagnetic instability with increasing interleg antiferromagnetic coupling. The boundary\nof the transition into the phase is obtained by studying the velocity r enormalization of the\ncorresponding gapless excitations. We mark the transition at uai= 0 (i= 1,2,3).\n3. Isotropic interleg exchange\nThe initial values of the coupling constants are g1(l= 0) =−2J⊥,g′\n1(l= 0) =−2J′\n⊥,\ng2(l= 0) = 2J⊥,g′\n2(l= 0) = 2J′\n⊥,g3(l= 0) = 2πJ⊥,g′\n3(l= 0) = 2πJ′\n⊥,g4(l= 0) =πJ⊥and\ng′\n4(l= 0) =πJ′\n⊥.\nFromtheRGequations(40)itisseenthatthemostrelevantoperat orsaretheg3,g′\n3terms.\nTherefore, the antisymmetric sector is gapped, and Θ aiare locked in the disordered phase.\nAs in the case of the transversal interleg interactions an effective sine-Gordon Hamiltonian\nfor the symmetric mode determines phase boundary ¯Ks= 1/2 between gapped and gapless\nphases. Numerical analysis shows that the ground state phase dia gram consists of the\ndisordered Rung Singlet gapfull phase and the stripe ferromagnet ic phase with dominating23\n/s48/s44/s48 /s48/s44/s53 /s49/s44/s48/s45/s49/s44/s48/s45/s48/s44/s56/s45/s48/s44/s54/s45/s48/s44/s52/s45/s48/s44/s50\n/s72/s97/s108/s100/s97/s110/s101/s32/s80/s104/s97/s115/s101\n/s83/s112/s105/s110/s32/s76/s105/s113/s117/s105/s100/s32/s73/s73/s32/s80/s104/s97/s115/s101/s32/s74\n/s122/s32/s39/s47/s74\n/s122/s32/s61/s32/s48/s46/s50\n/s74\n/s32/s47/s32/s74/s70/s101/s114/s114/s111/s109/s97/s103/s110/s101/s116/s32\n/s40/s97/s110/s116/s105/s112/s104/s97/s115/s101/s32/s105/s110/s116/s101/s114/s99/s104/s97/s105/s110/s32/s111/s114/s100/s101/s114/s41\n/s122\nFIG. 11: The ground-state phase diagram in the vicinity ∆ = −1 of the four-leg tube with\nlongitudinal coupling between legs.\nintraleg ferromagnetic ordering. The sector of the rung singlet ph ase increases at J′\n⊥→J⊥\n(see Fig. 12).\nA possible physical picture that reconciles the phase diagram with th e results of the\nprevious sections might look as follows. The system BIPNNBNO is an ar ray ofloosely\ncoupled ferrimagnetic chains in a presence of an extremely weak XY anisotropy and a strong\nfrustration, J′\n⊥∼J⊥, what corresponds to the region of the disordered Rung Singlet ph ase\nclose the point of the ferromagnetic instability, ∆ = −1.\nV. CONCLUSION\nWehave studied magnetizationprocess intwo-dimensional compoun dBIPNNBNO which\nexhibits ferrimagnetism of non-Lieb-Mattis type. The investigation is complicated by a lack\nof reliable information about exchange interactions in the system. F or a start, we proposed\nthenaive model ofnon-interacting ferrimagnetic chains andshowe d that anappearance both\n1/3 and 2/3 plateaus can be explained within the model. This provides u s the intrachain\nexchange couplings JAFandJ1. By setting these parameters in the exact diagonalization24\n/s48/s44/s48 /s48/s44/s49 /s48/s44/s50 /s48/s44/s51 /s48/s44/s52 /s48/s44/s53 /s48/s44/s54 /s48/s44/s55 /s48/s44/s56 /s48/s44/s57 /s49/s44/s48/s45/s49/s44/s48/s45/s48/s44/s57/s45/s48/s44/s56/s45/s48/s44/s55/s45/s48/s44/s54/s45/s48/s44/s53/s45/s48/s44/s52/s45/s48/s44/s51/s45/s48/s44/s50\n/s70/s101/s114/s114/s111/s109/s97/s103/s110/s101/s116\n/s40/s97/s110/s116/s105/s112/s104/s97/s115/s101/s32/s105/s110/s116/s101/s114/s99/s104/s97/s105/s110/s32/s111/s114/s100/s101/s114/s105/s110/s103/s41/s68/s105/s115/s111/s114/s100/s101/s114/s101/s100/s32\n/s82/s117/s110/s103/s32/s83/s105/s110/s103/s108/s101/s116/s115/s32\n/s74 /s32/s47/s74\nFIG. 12: The ground-state phase diagram in the vicinity ∆ = −1 of the four-leg tube with\nisotropic coupling between legs. The phase boundary betwee n the disordered rung singlet gapfull\nphase and the stripe ferromagnetic phase is shown by the dott ed and solid lines for J′\n⊥/J⊥= 0.2\nandJ′\n⊥/J⊥= 1.0, respectively.\nroutine the magnetization curve for the 32 and 40-sites clusters is numerically calculated.\nWe demonstrate that the magnetization curve similar to that obser ved in the experiment\nis obtained in the regime of weak interchain coupling, J1≫Jnn> Jnnn. Another revealed\nphenomenon is that a width of the singlet plateau increases with a gro wth of the antiferro-\nmagnetic frustrating coupling between the next-to-nearest cha ins. Following these results,\nwe apply on the tube lattice two low-energy theories which could expla in an appearance of\nthe singlet phase. The first one is based on the effective XXZ Heisenb erg model in a longitu-\ndinal magnetic field in the limit where the interchain coupling dominates, Jnn≥Jnnn≫J1.\nWe derive the critical field destroying the singlet plateau, it turns ou t that it does not\ndepend on the frustration parameter Jnnn. Another analytical strategy is realized via the\nabelian bosonization formalism which is relevant for the opposite limit J1≫Jnn≥Jnnn.\nWe demonstrate that the gapfull disordered Rung Singlets phase comes up when the XY\nexchange anisotropy may tilt the balance from the long-range orde r with an antiphase in-\nterchain arrangement of ferrimagnetic chains towards the spin liqu id phase. A role of the25\nanisotropy in a formation of the spin gap in the original two-dimension al system deserves a\nfurther study.\nAcknowledgments\nWe acknowledge Grant RFBR No. 10-02-00098-a for a support. We wish to thank Prof.\nD.N. Aristov for discussions.\nAppendix A: Wave functions of the ring\nThe states of the ring |S12S34SM/an}bracketri}htare composed from the functions |m1m2m3m4/an}bracketri}ht, where\nmi=±1/2 marks the spin-1/2 state of the separate i-th (1,1/2) block in the ring\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle11\n2;1\n2±1\n2/angbracketrightbigg\n=±/radicalbigg\n2\n3|1±1/an}bracketri}ht/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2∓1\n2/angbracketrightbigg\n∓1√\n3|10/an}bracketri}ht/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2±1\n2/angbracketrightbigg\n.\nThe basic functions of the singlet states are given by\n|00;00/an}bracketri}ht=1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2−1\n21\n2−1\n2/angbracketrightbigg\n−1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2−1\n2−1\n21\n2/angbracketrightbigg\n−1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n21\n21\n2−1\n2/angbracketrightbigg\n+1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n21\n2−1\n21\n2/angbracketrightbigg\n,\n|11;00/an}bracketri}ht=1√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2−1\n2−1\n2/angbracketrightbigg\n+1√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n2−1\n21\n21\n2/angbracketrightbigg\n−1\n2√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2−1\n21\n2−1\n2/angbracketrightbigg\n−1\n2√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2−1\n2−1\n21\n2/angbracketrightbigg\n−1\n2√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n21\n21\n2−1\n2/angbracketrightbigg\n−1\n2√\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n21\n2−1\n21\n2/angbracketrightbigg\n.\nThe triplet states read as\n|01;11/an}bracketri}ht=1√\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n2−1\n21\n21\n2/angbracketrightbigg\n−1√\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle−1\n21\n21\n21\n2/angbracketrightbigg\n,\n|10;11/an}bracketri}ht=1√\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n21\n2−1\n2/angbracketrightbigg\n−1√\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2−1\n21\n2/angbracketrightbigg\n.\nAppendix B: Saturation field in the limit of the strong ring coupling\nTo get the saturation field the functions of the ring with the total s pinsS= 5 andS= 6\n|ψ6/an}bracketri}ht=/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n1/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n4,\n|ψ5/an}bracketri}ht=−1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2/angbracketrightbigg\n1/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n4+1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n1/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2/angbracketrightbigg\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n426\n−1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n1/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2/angbracketrightbigg\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n4+1\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n1/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n2/vextendsingle/vextendsingle/vextendsingle/vextendsingle3\n23\n2/angbracketrightbigg\n3/vextendsingle/vextendsingle/vextendsingle/vextendsingle1\n21\n2/angbracketrightbigg\n4.\nwith the energies E5=JAF/2−Jnn/3+Jnnn/2,E6= 2JAF+Jnn+2Jnnnare needed.\nBy introducing the pseudo-spin operators in the restricted space similar to Eq.(23) the\noriginal spin operators are presented as follows ( S= 1,s= 1/2)\nSz\nin=23\n24+1\n12˜Sz\nn, sz\nin=5\n12+1\n6˜Sz\nn,\nS+\nin= (−1)i+11\n2/radicalbigg\n2\n3˜S+\nn, s+\nin= (−1)i1\n2/radicalbigg\n2\n3˜S+\nn,\nS−\nin= (−1)i+11\n2/radicalbigg\n2\n3˜S−\nn, s−\nin= (−1)i1\n2/radicalbigg\n2\n3˜S−\nn. (B1)\nThe effective XXZ spin chain Hamiltonian has the form (22) with the par ametersJeff\nxy=\n−2J1/3,Jeff\nz=J1/18 andBeff\nz=−7J1/9+B−Bc, whereBc=E6−E5.\nBy performing a Jordan-Wigner transformation followed by a partic le-hole\ntransformation20the new Hamiltonian of spinless holes is\nHh=−tN/summationdisplay\nn=1/parenleftbig\nd+\nidi+1+d+\ni+1di/parenrightbig\n+VN/summationdisplay\nn=1nd\nind\ni+1−µhN/summationdisplay\nn=1nd\ni,\nwheret=Jeff\nxy/2,V=Jeff\nzandµh=Jeff\nz−Beff\nz.\nThe saturation field Bsatcorresponds to the chemical potential where the hole band start s\nto fill up,µh=−2t. This yields Bsat=Bc+J1/6 = 3JAF/2+4Jnn/3+3Jnnn/2+J1/6.\n1F. Mila, Eur. J. Phys. 21, 499 (2000).\n2H. Kageyama, K. Yoshimura, R. Stern, N.V. Mushnikov, K. Oniz uka, M. Kato, K. Kosuge, C.P.\nSlichter, T. Goto, and Y. Ueda, Phys. Rev. Lett. 82, 3168 (1999).\n3S. Taniguchi, T. Nishikawa, Y. Yasui, Y. Kobayashi, M. Sato, T. Nishioka, M. Kotani, K. Sano,\nJ. Phys. Soc. Jpn. 64, 2758 (1995).\n4B.S. Shastry and B. Sutherland, Physica B 108, 1069 (1981).\n5C. Lacroix, P. Mendels, F. Mila (Eds.), Introduction to Frustrated Magnetism , (Springer, Berlin,\n2011).\n6J.S. Helton, K. Matan, M.P. Shores, E.A. Nytko, B.M. Bartlet t, Y. Yoshida, Y. Takano, A.\nSuslov, Y. Qiu, J.-H. Chung, D.G. Nocera, and Y. S. Lee, Phys. Rev. Lett. 98, 107204 (2007).27\n7P. Mendels, F. Bert, M.A. de Vries, A. Olariu, A. Harrison, F. Duc, J.C. Trombe, J.S. Lord, A.\nAmato, and C. Baines, Phys. Rev. Lett. 98, 077204 (2007).\n8S.-H. Lee, H. Kikuchi, Y. Qiu, B. Lake, Q. Huang, K. Habicht an d K. Kiefer, Nature Mater. 6,\n853 (2007).\n9F. Bert, D. Bono, P. Mendels, F. Ladieu, F. Duc, J.-C. Trombe, and P. Millet, Phys. Rev. Lett.\n95, 087203 (2005).\n10M. Yoshida, M. Takigawa, H. Yoshida, Y. Okamoto, and Z. Hiroi , Phys. Rev. Lett. 103, 077207\n(2009).\n11H. Nakano, T. Shimokawa, T. Sakai, will appear in J. Phys. Soc . Jpn (condmat/1110.3244v1).\n12T. Goto, N. V. Mushnikov, Y. Hosokoshi, K. Katoh, K. Inoue. Ph ysica B329-333 , 1160-1161\n(2003).\n13S. Brehmer, H.-J. Mikeska, and S. Yamamoto, J. Phys.: Conden s. Matter 9, 3921 (1997).\n14M. Oshikawa, M. Yamanaka, I. Affleck, Phys. Rev. Lett. 78, 1984 (1997).\n15T. Sakai, M. Sato, K. Okamoto, K. Okunishi and C. Itoi, J. Phys .: Condens. Matter 22, 403201\n(2010).\n16V.O. Garlea, A. Zheludev, L.-P. Regnault, J.-H. Chung, Y. Qi u, M. Boehm, K. Habicht, and\nM. Meissner, Phys. Rev. Lett. 100, 037206 (2008).\n17T. Sakai, and M. Takahashi, Phys. Rev. B 43, 13383 (1991); Phys. Rev. B 57, R3201 (1998).\n18V.E. Sinitsyn, I.G. Bostrem, A.S. Ovchinnikov. J. Phys. A: M ath. Theor. 40, 645-668 (2007).\n19I.G. Bostrem, V.E. Sinitsyn, A.S. Ovchinnikov, Y. Hosokosh i and K. Inoue, J. Phys.: Condens.\nMatter22, 036001 (2010).\n20F. Mila, Eur. Phys. J. B 6, 201 (1998).\n21G. Misguich andC. Lhuillier, in Frustrated spin systems , H.T. Diep, ed., World Scientific (2005).\n22P. Fulde, Electron Correlations in Molecules and Solids , (Springer-Verlag 1991), p.77.\n23P. Pfeuty, R. Jullien, and K.L. Penson in Real-Space Renormalization , T.W. Burkhardt and\nJ.M.J. van Leeuwen eds., (Springer, Berlin, 1982) Ch. 5.\n24E.H. Kim, J. S´ olyom, Phys. Rev. B 60, 15230 (1999).\n25D.C. Cabra, A. Honecker, and P. Pujol, Phys. Rev. B 58, 6241 (1998).\n26G. I. Japaridze, A. Langari and S. Mahdavifar, J. Phys.: Cond ens. Matter 19, 076201 (2007).\n27D.N. Aristov, M.N. Kiselev, Phys. Rev. B 70, 224402 (2004).\n28T. Kanzawa, Y. Hosokoshi, K. Katoh, S. Nishihara, K. Inoue, a nd H. Nojiri, J. Phys.: Conf.28\nSer.51, 91 (2006).\n29M. Oshikawa and I. Affleck, Phys. Rev. B 65, 134410 (2002).\n30A.K. Kolezhuk and H.-J. Mikeska, Phys. Rev. B 53, R8848 (1996).\n31M. Roji and S. Miyashita, J. Phys. Soc. Jpn. 65, 883 (1996).\n32R.M. Wießner, A. Fledderjohann, K.-H. M¨ uller, and M. Karba ch, Phys. Rev. B 60, 6545 (1999).\n33A. Luther and I. Peschel, Phys. Rev. B 12, 3908 (1975).\n34T. Giamarchi, Quantum Physics in One Dimension (Oxford University Press, Oxford, 2003).\n35A.M. Tsvelik, Quantum Field Theory in Condensed Matter Physics (Cambridge University\nPress, Cambridge, 1995).\n36H.J. Schulz, Phys. Rev. B 34, 6372 (1986).\n37T. Vekua, G.I. Japaridze, and H.-J. Mikeska, Phys. Rev. B 67, 064419 (2003)." }, { "title": "1304.6525v1.Observation_of_an_inter_sublattice_exchange_magnon_in_CoCr__2_O__4__and_analysis_of_magnetic_ordering.pdf", "content": "arXiv:1304.6525v1 [cond-mat.mtrl-sci] 24 Apr 2013Observation of an inter-sublattice exchange magnon in CoCr 2O4\nand analysis of magnetic ordering\nD. Kamenskyi,1H. Engelkamp,2T. Fischer,1M. Uhlarz,1J. Wosnitza,1B. P. Gorshunov,3,4,5\nG. A. Komandin,3A. S. Prokhorov,3,4M. Dressel,5A. A. Bush,6V. I. Torgashev,7and A. V. Pronin1,∗\n1Dresden High Magnetic Field Laboratory (HLD),\nHelmholtz-Zentrum Dresden-Rossendorf, 01314 Dresden, Ger many\n2High Field Magnet Laboratory, Institute for Molecules and M aterials,\nRadboud University Nijmegen, 6525 ED Nijmegen, The Netherl ands\n3A. M. Prokhorov Institute of General Physics, Russian Acade my of Sciences, 119991 Moscow, Russia\n4Moscow Institute of Physics and Technology (State Universi ty), 141700 Dolgoprudny, Moscow Region, Russia\n51. Physikalisches Institut, Universit¨ at Stuttgart, Pfaffe nwaldring 57, 70550 Stuttgart, Germany\n6Moscow State Institute of Radio-Engineering, Electronics ,\nand Automation (Technical University), 117464 Moscow, Rus sia\n7Faculty of Physics, Southern Federal University, 344090 Ro stov-on-Don, Russia\n(Dated: May 24, 2022)\nWe report on an investigation of optical properties of multi ferroic CoCr 2O4at terahertz frequen-\ncies in magnetic fields up to 30 T. Below the ferrimagnetic tra nsition (94 K), the terahertz response\nof CoCr 2O4is dominated by a magnon mode, which shows a steep magnetic-fi eld dependence. We\nascribe this mode to an exchange resonance between two magne tic sublattices with different g-\nfactors. In the framework of a simple two-sublattice model ( the sublattices are formed by Co2+\nand Cr3+ions), we find the inter-sublattice coupling constant, λ=−(18±1) K, and trace the\nmagnetization for each sublattice as a function of field. We s how that the Curie temperature of the\nCr3+sublattice, Θ 2= (49±2) K, coincides with the temperature range, where anomalies of the\ndielectric and magnetic properties of CoCr 2O4have been reported in literature.\nPACS numbers: 75.85.+t, 76.50.+g\nCoCr2O4is a ferrimagnetic spinel compound with a\ncomplex network of competing magnetic interactions.1,2\nBoth, Co2+(A sites of the spinel structure) and Cr3+\nions (B1 and B2 sites), are magnetic. Below the Curie\ntemperature, TC= 94 K, the system exhibits a long-\nrange ferrimagnetic order.1AtTS= 26 K, a structural\ntransition occurs. Below this temperature, an incom-\nmensurate conical (i.e. uniform plus transverse spiral)\nmagnetic structure sets in. At Tlock−in= 15 K, the mag-\nnetic structure becomes commensurate – the period of\nthe spin spiral “locks” to the lattice parameter.2\nThe spiral order most likely survives above TS, but\non short ranges only.1,2AtTkink= 50 K, anomalies in\ndielectric and magnetic properties of CoCr 2O4have been\nreported3,4and tentativelyattributed to the formationof\nthis incommensurate short-range spiral magnetic order.4\nIn 2006, Yamasaki et al.have discovered the struc-\ntural transition at TS= 26 K to be accompanied by\nthe emergence of a spontaneous electric polarization,\nwhich direction can be reversed by applying a magnetic\nfield.5As multiferroics are appealing because of basic\nphysical interest as well as their potential technological\napplications,6the reported multiferroicity of CoCr 2O4\nhas triggered a number of experimental studies of the\ncompound.4,7–10To date, there is, however, no optical\ndata reported at terahertz frequencies. This is the fre-\nquency region where e.g.electromagnons have been ob-\nserved in some other multiferroic systems.11,12\nHere, we report on an optical study of CoCr 2O4at ter-\nahertz (or far-infrared) frequencies in magnetic fields upto 30 T. We do not find any evidence for electromagnons.\nHowever, below TC, we observe a resonance mode, which\nis highly sensitive to external magnetic field. We ascribe\nthis mode to a ferrimagnetic inter-sublattice exchange\nresonance, originally considered theoretically by Kaplan\nand Kittel.13Applying this model to CoCr 2O4allows us\nto separate the contributions of the Co2+and Cr3+sub-\nlattices to the total magnetization, to extract the mag-\nnetization for each sublattice as a function of field, and\nto find the inter-sublattice coupling constant. Further-\nmore, we show that the Curie temperature of the Cr3+\nsublattice coincides with Tkink= 50 K. Thus, the onset\nof the short-range spiral component must be related to\nthe ordering in the Cr3+sublattice.\nThe investigated CoCr 2O4samples have been synthe-\nsized from Co 3O4and Cr 2O3powders. CoCr 2O4pow-\nder has been pressed into pellets of 10 mm in diameter\nand 1 mm in thickness. X-ray diffraction measurements\nprove the cubic symmetry (space group Fd¯3m) with\nno indication of spurious phases. The lattice constant,\na= 8.328(2)˚A, is in good agreement with published\nresults.14,15More details on sample preparation, char-\nacterization, and thermodynamic properties of CoCr 2O4\nare given in our recent works.3,16\nThe optical measurements have been made in trans-\nmission mode, using two different setups. The first setup\nwas a spectrometer equipped with backward-wave os-\ncillators (BWOs) as sources of coherent and frequency-\ntunable radiation.17,18The measurements were done at\n(sub)terahertz frequencies ( ν=ω/2π= 8.5÷41 cm−1=2\n/s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48/s76/s105/s110/s101/s32/s50/s32/s63/s84/s114/s97/s110/s115/s109/s105/s116/s116/s97/s110/s99/s101/s32/s40/s97/s114/s98/s46/s32/s117/s110/s105/s116/s115/s41/s51/s32/s84\n/s53/s32/s84\n/s56/s32/s84\n/s49/s48/s32/s84\n/s49/s50/s32/s84\n/s49/s53/s32/s84\n/s49/s55/s32/s84\n/s50/s48/s32/s84\n/s50/s53/s32/s84\n/s51/s48/s32/s84\n/s32 /s32\n/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32/s40/s99/s109/s45/s49\n/s41/s67/s111/s67/s114\n/s50/s79\n/s52\n/s76/s105/s110/s101/s32/s49\n/s51/s48/s32/s75\nFIG. 1: (Color online) Examples of raw far-infrared transmi t-\ntance spectra of CoCr 2O4in Faraday geometry. The absorp-\ntion line, marked as “Line 1”, is discussed in the course of th e\narticle. At the higher-frequency end of the spectra, there i s\nan indication of another, very broad, line (“Line 2 ?”).\n250÷1230 GHz = 1 ÷5 meV) and at temperatures be-\ntween 4 and 300 K. The radiation was linearly polarized.\nThe absolute values of optical transmission (normalized\nto the empty-channel measurements) were obtained in\nessentially the same way as described, e.g., in Ref. 19.\nA commercial split-coil magnet, embedded into an op-\ntical cryostat, was utilized for measurements in magnetic\nfields up to 8 T. In these measurements, we used three\ndifferent geometries: k⊥H/bardbl˜h,k⊥H⊥˜h, and\nk/bardblH⊥˜h(hereHis the external static magnetic field;\nkand˜harethe wavevectorand the magnetic component\nof the probing electromagnetic radiation, respectively).\nIn the first geometry ( k⊥H/bardbl˜h), where electro-\nmagnons, if they exist, could be excited, we did not ob-\nserve any resonance absorption lines. Furthermore, these\nspectra show no detectable changes induced by applying\nmagnetic fields up to 8 T.\nIn the two other geometries (where H⊥˜h), we ob-\nserved absorption lines. We did not detect any differ-\nences between the results for these two geometries, al-\nthough the data obtained for k/bardblH(Faraday geometry)\nare somewhat less noisy due to peculiarities of the split-\ncoil magnet design. Thus, in the following we discuss the\nBWO data collected in the Faraday geometry.\nThe second optical setup was a commercial Fourier-\ntransform infrared (FTIR) spectrometer (Bruker\nIFS113v) combined with a continuous-field 33-Tesla\nBitter magnet, at the High Field Magnet Laboratory in\nNijmegen.20The measurements were performed in the\nFaradaygeometry( k/bardblH⊥˜h) at 4 and 30K. A mercury\nlamp was used as a radiation source. The far-infrared/s49/s48/s50/s48/s51/s48/s52/s48/s53/s48\n/s48 /s49/s48 /s50/s48 /s51/s48/s49/s48/s50/s48/s51/s48/s52/s48/s53/s48\n/s32\n/s32/s66/s87 /s79/s32/s109/s101/s97/s115/s117/s114/s101/s109/s101/s110/s116/s115\n/s32/s70/s84/s73/s82/s32/s109/s101/s97/s115/s117/s114/s101/s109/s101/s110/s116/s115\n/s32/s77/s111/s100/s101/s108/s32/s99/s97/s108/s99/s117/s108/s97/s116/s105/s111/s110/s115/s32/s32\n/s51/s48/s32/s75/s32/s82/s101/s115/s111/s110/s97/s110/s99/s101/s32/s102/s114/s101/s113/s117/s101/s110/s99/s121/s32\n/s48/s32/s40/s99/s109/s45/s49\n/s41/s67/s111/s67/s114\n/s50/s79\n/s52\n/s52/s32/s75/s32\n/s48/s72/s32/s40/s84/s41\n/s32/s32\nFIG. 2: (Color online) Field dependence of the magnon fre-\nquency for 4 and 30 K. Symbols are the experimental data,\nlines represent the results of simultaneous fits of the optic al\ndata with Eq. 3 and of magnetization data from Ref. 3 with\nEqs. 1 and 2. Note that the onset of the vertical scales is at\n10 cm−1.\nradiation was detected using a custom-made silicon\nbolometer operating at 1.4 K. For both temperatures,\nthe FTIR spectra were measured in 0, 3, 5, 8, 10,\n12, 15, 17, 20, 25, and 30 T. The optical data were\ncollected between 15 and 70 cm−1(450÷2100 GHz,\n1.9÷8.7 meV), using a 200- µm mylar beamsplitter and\na scanning velocity of 50 kHz. At each field, at least 100\nscans were averaged.\nIn the transmission spectra, obtained by use of both\nsetups, we reliably observed an intensive absorption line,\nwhich is highly sensitive to the applied magnetic field\nand temperature (“Line 1” in Fig. 1). At the higher-\nfrequency end of the FTIR spectra, we possibly observe\nanother, broad and week, absorption line (“Line 2 ?” in\nFig. 1). The position of this line shifts to higher frequen-\ncies (and thus out of the measurement-frequency win-\ndow) in applied magnetic field. We tentatively attribute\nthis line to an antiferromagnetic resonance within the\nCr2+sublattice. Hereafter, we solely discuss the “Line\n1”.\nThe field dependence of the frequency of “Line 1”, ν0,\nis shown in Fig. 2. The resonance frequency is basically\nproportional to the applied static field with a zero-field\ngap. At 4 K, the gap is 15.8 cm−1(475 GHz, 1.97 meV),\nand at 30 K, it is 16.3cm−1(489GHz, 2.02meV). At low\nfields[µ0H<∼10T],thedimensionlessslopesofthe ν0(H)3\ncurves,hν0/µBµ0H, reach 2.5(at 4 K) and 2.8 (at 30K).\nLet us note, that the typical slopes of the ν0(H) curves\nfor ferromagnetic resonances in the Co2+- or Cr3+-based\ncompounds are determined by the gyromagnetic ratios\nof the Co2+and Cr3+ions and amount to 2.2 and 1.95,\nrespectively.21\nResonance modes in ferrimagnetic substances have\noriginally been considered theoretically by Kaplan and\nKittel.13For a system with two magnetic sublattices,\nthey predicted the existence of two magnetic-resonance\nlines. The first of them is associated with the conven-\ntional spin precession. In CoCr 2O4, such mode has been\nobserved by electron-spin-resonance studies at frequen-\ncies below 100 GHz in fields up to 10 T.22,23The sec-\nond mode, the inter-sublattice exchange resonance, is\nsupposed to have a large zero-field gap defined by the\nexchange interaction between the sublattices, to harden\nin an applied magnetic field, and to have a steep field\ndependence.24Our observed mode demonstrates this\nvery behavior (Fig. 2). Thus, it is reasonable to asso-\nciate it with the Kaplan-Kittel inter-sublattice exchange\nresonance.\nIn the following, we show that using a simple two-\nsublattice Kaplan-Kittel model, we can consistently de-\nscribe our data on the magnon frequency and our earlier\ndata on magnetization (Ref. 3). From this description,\nwe obtain the magnetic moments of the sublattices and\nthe inter-sublattice exchange constant.\nIn our model, we consider two effective magnetic sub-\nlattices. Coupling within the first sublattice is provided\nby exchange interactions between Co2+and Cr3+ions,\nJAB. At Θ 1=TC= 94 K, this sublattice orders. Im-\nportant is that only non-collinearly ordered spins of the\nCo2+ions contribute to the magnetic moment, while the\nspins of the Cr3+ions do not. This is because of geo-\nmetrical frustration of the interaction between the Cr3+\nions,JBB.2Thus, one can consider the first sublattice to\nbe formed by Co2+ions only.\nWhen temperature decreases further, the Cr3+spins\nstart to contribute (negatively) to the net magnetiza-\ntion. Thus, as the second sublattice, one can consider\nthe sublattice formed by Cr3+ions. Let us note, that\nbecause of geometrical frustration, the effective coupling\nwithin the Cr3+sublattice is significantly smaller than\nJBB. The Curie temperature of this sublattice, Θ 2, is\nsomewhere below 94 K. This temperature will be one of\nour fit parameters (we take Θ 2to be field independent).\nFor a two-sublattice ferrimagnet, the net (measurable)\nmagnetization is the sum of the magnetizations of the\nsublattices, M(T,H) =M1(T,H)+M2(T,H).M1and\nM2have opposite signs; the inter-sublattice exchange\nconstant λis negative. Within the molecular-field ap-\nproximation, the reduced magnetization of each sublat-\ntice,yi≡Mi/M0\ni[here and thereafter i={1,2} ≡\n{Co2+,Cr3+}andM0\niis the zero-temperature magne-\ntization of the i-th sublattice], is given by the Brillouin/s48 /s52/s48 /s56/s48 /s49/s50/s48/s45/s48/s46/s52/s45/s48/s46/s50/s48/s46/s48/s48/s46/s50/s48/s46/s52\n/s119/s111/s117/s108/s100/s45/s98/s101/s32/s99/s117/s114/s118 /s101/s115/s58\n/s32/s67/s114/s51/s43\n/s32/s119/s105/s116/s104/s111/s117/s116/s32/s99/s111/s117/s112/s108/s105/s110/s103\n/s98/s101/s116/s119/s101/s101/s110/s32/s116/s104/s101/s32/s115/s117/s98/s108/s97/s116/s116/s105/s99/s101/s115\n/s32/s67/s114/s51/s43\n/s32/s119/s105/s116/s104/s111/s117/s116/s32/s116/s97/s107/s105/s110/s103\n/s105/s110/s116/s111/s32/s97/s99/s99/s111/s117/s110/s116/s32/s116/s104/s101/s32/s115/s116/s114/s117/s99/s116/s117/s114/s97/s108\n/s116/s114/s97/s110/s115/s105/s116/s105/s111/s110/s32/s97/s116/s32/s50/s54/s32/s75/s115/s117/s98/s108/s97/s116/s116/s105/s99/s101\n/s99/s111/s110/s116/s114/s105/s98/s117/s116/s105/s111/s110/s115/s58\n/s32/s67/s111/s50/s43\n/s32/s67/s114/s51/s43/s32/s32/s101/s120/s112/s101/s114/s105/s109/s101/s110/s116\n/s32/s32/s116/s111/s116/s97/s108/s32/s102/s105/s116\n/s32/s32/s77/s97/s103/s110/s101/s116/s105/s122/s97/s116/s105/s111/s110 /s32/s40\n/s66 /s32/s47/s32/s102/s46/s117/s46 /s41\n/s84/s101/s109/s112/s101/s114/s97/s116/s117/s114/s101/s32/s40/s75/s41/s48/s72/s32/s61/s32/s48/s46/s49/s32/s84/s67/s111/s67/s114\n/s50/s79\n/s52\nFIG. 3: (Color online) Example of a fit by use of Eqs. 1 and 2\nto the magnetization data obtained at 0.1 T [simultaneously ,\nthe magnon-frequency data were fitted with the same set of\nfree parameters, see Fig. 2]. Contributions of the Co2+and\nCr3+sublattices are shown with opposite signs. The contri-\nbution of the Cr3+sublattice diminishes at TC, rather than at\nΘ2= 49 K, because of the coupling between the sublattices.\nfunction:25,26\nyi=BSi(xi)≡2Si+1\n2Sicoth2Si+1\n2Sixi−1\n2Sicothxi\n2Si,\n(1)\nwhereSiis the spin magnetic moment for ions of the i-th\nsublattice and xiis defined as:\nxi=µB\nkBT/bracketleftbigg3Si\nSi+1/parenleftbiggkBΘiMi\nµBM0\ni+kBλMj\nµBM0\ni/parenrightbigg\n+HM0\ni/bracketrightbigg\n,(2)\ni/ne}ationslash=j.\nIn our case, Co2+and Cr3+ions have equal spin mo-\nments,S1=S2= 3/2.\nFor the resonance frequency ν0, we use a modified\nKaplan-Kittel equation:10,13\n/parenleftbigghν0\nµBg1+kB\nµBλM2+H/parenrightbigg\n×(3)\n/parenleftbigghν0\nµBg2−kB\nµBλM1+H/parenrightbigg\n+/parenleftbiggkB\nµBλ/parenrightbigg2\nM1M2= 0,\nwhereg1= 2.2 and g2= 1.95 are the gyromagneticratios\nfor the Co2+and Cr3+ions, respectively.21In Eqs. 2 and\n3, magnetization is in dimensionless units. In our model,\nwe neglect anisotropy, as our measurements have been\nperformed on powder samples.274\n/s48 /s49/s48 /s50/s48 /s51/s48/s45/s48/s46/s52/s45/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54/s48/s46/s55\n/s67/s111/s67/s114\n/s50/s79\n/s52\n/s48/s72/s32/s40/s84/s41/s84/s32/s61/s32/s51/s48/s32/s75\n/s32/s32/s83/s117/s98/s108/s97/s116/s116/s105/s99/s101/s32/s109/s97/s103/s110/s101/s116/s105/s122/s97/s116/s105/s111/s110/s32/s40 /s32/s47/s32/s102/s46/s117/s46/s41/s67/s111/s50/s43\n/s67/s114/s51/s43\nFIG. 4: (Color online) Magnetization of the Co2+and Cr3+\nsublattices in CoCr 2O4, as obtained from the model fits with\nEqs. 1 – 3. The lowest-field points are at 0.1, 0.5, and 1 T.\nUnlike in collinear ferrimagnetics, in CoCr 2O4the vec-\ntor of the magnetization of each sublattice in an exter-\nnal magnetic field can change not only its orientation,\nbut also its size. In our calculations, we take this into\naccount. The corresponding increase in M0\n1andM0\n2is\nsupposed to persist until the spin-only values for Co2+\nand Cr3+ions of 3 µB/ion are reached. Recent experi-\nments in pulsed magnetic fields have shown that even in\nfields of up to 60 T, there are no signs of saturation in\nthe magnetization.28\nUsing Eqs. 1 – 3, we simultaneously fit the magneti-\nzation data, obtained in Ref. 3, and our results for the\nresonance frequency.29Results of these fits are shown in\nFigs. 2 and 3 together with experimental data. We find\nthat the best description of all our experimental data can\nbe reached with λ=−(18±1) K and Θ 2= (49±2) K.\nThe value found for the ordering temperature of the\nCr3+sublattice, Θ 2, coincides with Tkink– the tem-\nperature, which is possibly related to the formation of\nthe incommensurate short-range spiral magnetic order.4\nOur findings show that this short-range order appears to\nbe due to the involvement of the Cr3+ions. The rela-\ntively large value of the coupling constant, λ=−18 K,\nmay explain why the Tkink-related features appear to be\nbroad.2,4Figure 4 shows the magnetic-field dependence of the\ncalculated sublattice magnetizations. We show here the\nresults obtained at T= 30 K, i.e. just above the struc-\ntural transition. As it can be seen from Fig. 3, the\nlower-temperatureresults differ only marginallyfrom the\nshown in Fig. 4. We have found that within our model,\nthe observed behavior of the experimental data at TS\ncan be best described by a slight decrease in the absolute\nvalue ofMCratTS(theg-factors of the Co2+and Cr3+\nions are presumed to be independent of temperature).\nThe contribution of each sublattice to the total mag-\nnetic moment in zero field can be estimated from Fig. 4.\nWe findMCo(0 T) = 0.4 µB/f.u. and MCr(0 T) = – 0.33\nµB/f.u. We note, that although in fields smaller than\nthe coercive force (0.3 T, Ref. 3) the Mivalues are influ-\nenced by the magnetizing/demagnetizing processes, this\ninfluence hardly affects the above result [in Fig. 4, only\nthe lowest-field point (0.1 T) is below the coercive-force\nvalue, the next points are at 0.5 and 1 T].\nThe proposed model gives a natural explanation for\nthe steep slope of the ν0(H) curves [Fig. 2]. As can be\nseen from Eq. 3, this is due to the increase of Mias\na function of field. Similarly, the temperature evolution\nofν0(H) is explained as being due to the temperature\nevolution of the Mi(H) curves.\nSummarizing, the far-infrared optical response of\nCoCr2O4below the Curie temperature, TC= 94 K, is\ndominated by a magnon mode. Ascribing the magnon to\nan inter-sublattice exchange resonance (the sublattices\nare formed by the Co2+and Cr3+ions) and applying a\nmodified Kaplan-Kittel model of two interacting sublat-\ntices allows us to consistently describe our experimental\ndata on the magnon frequency and earlier magnetization\nmeasurements, to obtain the magnetization for each of\nthe sublattices, and to find the inter-sublattice exchange\nconstant, λ= – 18 K. The Curie temperature of the\nCr3+sublattice is found to coincide with the tempera-\nture, where the incommensurate short-range spiral mag-\nnetic order is believed to set in, Tkink= 50 K.\nWe thank M. D. Kuz’min for useful discussions and\nPapori Gogoi for assistance in measurements. Parts of\nthis work were supported by EuroMagNET II (EU con-\ntract No. 228043), by the Russian Foundation for Basic\nResearch (Grant No. 12-02-00151) and by the Ministry\nofEducation and Science ofthe Russian Federation (con-\ntract No. 14.A18.21.0740).\n∗Electronic address: a.pronin@hzdr.de\n1N. Menyuk, K. Dwight, and A. Wold, J. Phys. (Paris) 25,\n528 (1964).\n2K. Tomiyasu, J. Fukunaga, and H. Suzuki, Phys. Rev. B\n70, 214434 (2004).\n3A.V.Pronin, M.Uhlarz, R.Beyer, T.Fischer, J.Wosnitza,\nB. P. Gorshunov, G. A. Komandin, A. S. Prokhorov, M.\nDressel, A. A. Bush, and V. I. Torgashev, Phys. Rev. B85, 012101 (2012).\n4G. Lawes, B. Melot, K. Page, C. Ederer, M. A. Hayward,\nTh. Proffen, and R. Seshadri, Phys. Rev. B 74, 024413\n(2006).\n5Y. Yamasaki, S. Miyasaka, Y. Kaneko, J.-P. He, T. Arima,\nand Y. Tokura, Phys. Rev. Lett. 96, 207204 (2006).\n6T. Kimura, T. Goto, H. Shintan, K. Ishizaka, T. Arima,\nand Y. Tokura, Nature 426, 55 (2003).5\n7Y. J. Choi, J. Okamoto, D. J. Huang, K. S. Chao, H. J.\nLin, C. T. Chen, M. van Veenendaal, T. A. Kaplan, and\nS.-W. Cheong, Phys. Rev. Lett. 102, 067601 (2009).\n8N. Mufti, A. A. Nugroho, G. R. Blake, and T. T. M. Pal-\nstra, J. Phys.: Condens. Matter 22, 075902 (2010).\n9L. J. Chang, D. J. Huang, W.-H. Li, S.-W. Cheong, W.\nRatcliff, and J. W. Lynn, J. Phys.: Condens. Matter 21,\n456008 (2010).\n10V. I. Torgashev, A. S. Prokhorov, G. A. Komandin, E. S.\nZhukova, V. B. Anzin, V. M. Talanov, L. M. Rabkin, A. A.\nBush, M. Dressel, and B. P. Gorshunov, Phys. Solid State\n54, 350 (2012).\n11A. Pimenov, A. A. Mukhin, V. Y. Ivanov, V. D. Travkin,\nA. M. Balbashov, and A. Loidl, Nat. Phys. 2, 97 (2006).\n12A. B. Sushkov, R. V. Aguilar, S.-W. Cheong, and H. D.\nDrew, Phys. Rev. Lett. 98, 027202 (2007).\n13J. Kaplan and C. Kittel, J. Chem. Phys. 21, 760 (1953).\n14B. Mansour, N. Baffier, and M. Huber, C. R. Hebd.\nSeances Acad. Sci., Ser. C 277, 867 (1973).\n15P. G. Casado and I. Rasines, Polyhedron 5, 787 (1986).\n16M. Uhlarz, A. V. Pronin, J. Wosnitza, A. S. Prokhorov,\nand A. A. Bush, Phys. Chem. Minerals 40, 203 (2013).\n17G. V. Kozlov and A. A. Volkov, in Millimeter and Submil-\nlimeter Wave Spectroscopy of Solids , edited by G. Gr¨ uner\n(Springer, Berlin, 1998), p. 51.\n18B. Gorshunov, A. Volkov, I. Spektor, A. Prokhorov, A.\nMukhin, M. Dressel, S. Uchida, and A. Loidl, Int. J. In-\nfrared Millim. Waves 26, 1217 (2005).\n19A. Mukhin, B. Gorshunov, M. Dressel, C. Sangregorio, andD. Gatteschi, Phys. Rev. B 63, 214411 (2001).\n20S. A. J. Wiegers, P. C. M. Christianen, H. Engelkamp, A.\nden Ouden, J. A. A. J. Perenboom, U. Zeitler, and J. C.\nMaan, J. Low Temp. Phys. 159, 389 (2010).\n21S. A. Altshuler and B. M. Kozyrev, Electron Paramagnetic\nResonance in Compounds of Transition Elements (Wiley,\nNew York, 1974).\n22J. J. Stickler and H. J. Zeiger, J. Appl. Phys. 39, 1021\n(1968).\n23S. Funahashi, K. Siratori, and Y. Tomono, J. Phys. Soc.\nJpn.29, 1179 (1970).\n24T. A. Kaplan, Phys. Rev. 109, 782 (1958).\n25S. V. Vonsovsky, Magnetism (Wiley, New York, 1974).\n26M. D. Kuz’min and A. M. Tishin, Cryogenics 32, 545\n(1992).\n27In practice, this approximation would also be valid for sin-\ngle crystals, as the magnitude of the anisotropy field in\nCoCr2O4is below 0.1 T [Ref. 23], i.e. significantly smaller\nthan all terms in Eqs. 2 and 3 for external fields above 0.1\nT.\n28V. Tsurkan, S. Zherlitsyn, S. Yasin, V. Felea, Y. Skourski,\nJ. Deisenhofer, H.-A. Krug von Nidda, J. Wosnitza, and\nA. Loidl, Phys. Rev. Lett. 110, 115502 (2013).\n29Since no experimental data on the temperature depen-\ndence of the magnetization at 30 T are available, we used\na linear extrapolation of the magnetization in accordance\nwith the M(H) measurements of Ref. 28." }, { "title": "1207.6771v1.Ferrimagnetism_of_the_Heisenberg_Models_on_the_Quasi_One_Dimensional_Kagome_Strip_Lattices.pdf", "content": "arXiv:1207.6771v1 [cond-mat.str-el] 29 Jul 2012Ferrimagnetism of the Heisenberg Models on the Quasi-One-D imensional Kagome\nStrip Lattices\nTokuro Shimokawa1and Hiroki Nakano1\n1Graduate School of Material Science, University of Hyogo, K amigori, Hyogo 678-1297, Japan\n(Dated: May 27, 2018)\nWe study the ground-state properties of the S= 1/2 Heisenberg models on the quasi-one-\ndimensional kagome strip lattices by the exact diagonaliza tion and density matrix renormalization\ngroup methods. The models with two different strip widths sha re the same lattice structure in their\ninner part with the spatially anisotropic two-dimensional kagome lattice. When there is no mag-\nnetic frustration, the well-known Lieb-Mattis ferrimagne tic state is realized in both models. When\nthe strength of magnetic frustration is increased, on the ot her hand, the Lieb-Mattis-type ferrimag-\nnetism is collapsed. We find that there exists a non-Lieb-Mat tis ferrimagnetic state between the\nLieb-Mattis ferrimagnetic state and the nonmagnetic groun d state. The local magnetization clearly\nshows an incommensurate modulation with long-distance per iodicity in the non-Lieb-Mattis ferri-\nmagnetic state. The intermediate non-Lieb-Mattis ferrima gnetic state occurs irrespective of strip\nwidth, which suggests that the intermediate phase of the two -dimensional kagome lattice is also the\nnon-Lieb-Mattis-type ferrimagnetism.\nPACS numbers: 75.10.Jm, 75.30.Kz, 75.40.Mg\nI. INTRODUCTION\nFerrimagnetism is a fundamental phenomenon in the\nfield of magnetism. The simplest case of the conven-\ntional ferrimagnetism is the ground state of the mixed\nspin chain. For example, there is an ( s,S)=(1/2, 1)\nmixed spin chain with a nearest-neighbor antiferromag-\nnetic (AF) isotropic interaction. In this system, the\nso-called Lieb-Mattis (LM)-type ferrimagnetism1–6is re-\nalized in the ground state because two different spins\nare arranged alternately in a line owing to the AF in-\nteraction. Since this type of ferrimagnetism has been\nstudied extensively, the magnetic properties and the oc-\ncurrence mechanism of the LM ferrimagnetism are well\nknown. In particular, the ferrimagnetism in the quan-\ntum Heisenberg spin model on the bipartite lattice with-\nout frustration is well understood within the Marshall-\nLieb-Mattis (MLM) theorem1,2. In the LM ferrimagnetic\nground state, the spontaneous magnetization occurs and\nthe magnitude is fixed to a simple fraction of the satu-\nrated magnetization.\nIn recent years, a new type of ferrimagnetism, which is\nclearlydifferent fromLM ferrimagnetism, has been found\nin the ground state of several one-dimensional frustrated\nHeisenberg spin systems7–13. The spontaneous magneti-\nzation in this new type of ferrimagnetism changes grad-\nually with respect to the strength of frustration. The\nincommensurate modulation with long-distance period-\nicity in local magnetizations is also a characteristic be-\nhavior of the new type of ferrimagnetism. Hereafter, we\ncall the new type of ferrimagnetism the non-Lieb-Mattis\n(NLM) type. The mechanism of the occurrence of the\nNLM ferrimagnetism has not yet been clarified.\nOn the other hand, some candidates of the NLM ferri-\nmagnetism among the two-dimensional (2D) systems, for\nexample, the mixed-spin J1-J2Heisenberg model on the\nsquarelattice14andtheS= 1/2HeisenbergmodelontheUnion Jack lattice15, were reported. These 2D frustrated\nsystems have the intermediate ground state, namely, the\n“canted state”, in which the spontaneous magnetization\nis changed when the inner interaction of the system is\nvaried. It has not been, however, determined whether\nthe incommensurate modulation with long-distance pe-\nriodicity exists in the local magnetization of the canted\nstate owing to the difficulty of treating these 2D frus-\ntrated systems numerically and theoretically. Therefore,\nthe relationships between the canted states of these 2D\nfrustrated systems and the NLM ferrimagnetic state are\nstill unclear.\nUnder these circumstances, recently, another candi-\ndate of the NLM ferrimagnetism among the 2D systems\nwas reported in ref. VI in which the S= 1/2 Heisen-\nberg model on the spatially anisotropic kagome lattice\ndepicted in Fig. 1(a) was studied. A region of the\nintermediate-magnetization states is observed between\nthe LM ferrimagnetism that is rigorously proved by the\nMLM theorem1,2and the nonmagnetic state of the spa-\ntially isotropic kagome-lattice antiferromagnet in the ab-\nsence of magnetic field17–27. The local magnetization in\nthe intermediate state of the kagome lattice was investi-\ngated by the exact diagonalization method, and it was\nreported that the local magnetization greatly depends\non the position of the sites, although it is difficult to\ndetermine clearly whether the incommensurate modula-\ntion with long-distance periodicity is present. This result\nleads to the high expectation that the intermediate state\nof the spatially anisotropic kagome lattice is the NLM\nferrimagnetic state. Additional research is desirable to\nconclude that the intermediate state of this 2D system is\nthe NLM ferrimagnetism.\nIn this paper, we study the ground-state properties\nof theS= 1/2 Heisenberg models on the quasi-one-\ndimensional (Q1D) kagome strip lattices depicted in\nFigs. 1(b) and 1(c) instead of the 2D lattice depicted2\n+'\n+\u0013+\u0012\n+\u0012+\u0013\n\"` \"\n#\n$\n%\n\"` #` \n$` \"\n# +\u0012\n+\u0013+'\n$\tB\n \tC\n \tD\nFIG. 1: Structures of the lattices: the spatially anisotrop ic kagome lattice (a) and the quasi-one-dimensional kagome strip\nlattices (b) and (c) with different widths. An S= 1/2 spin is located at each site denoted by a black circle. Antif erromagnetic\nbondsJ1(bold straight line) and J2(dashed line), and ferromagnetic bond JF(dotted line). The sublattices in a unit cell of\nlattice (b) are represented by A, A′, B, B′, C, C′, and D. The sublattices in a unit cell of lattice (c) are repre sented by A, A′,\nB, and C.\nin Fig. 1(a). Note that the inner parts of the lattices in\nFigs. 1(b) and 1(c) are common to a part of the 2D lat-\ntice in Fig. 1(a). We also note that the lattice shapes of\nstripsinthepresentstudyaredifferentfromsomekagome\nstrips (chains) studied in refs. VI-VI, where the nontriv-\nial properties of kagome antiferromagnets were reported.\nAccordingtothestudyinref. VI,itwasalreadyknown\nthat the NLM ferrimagnetism is realized in the ground\nstate of the kagome strip lattice in Fig. 1(c). In the\npresent study, we show that both the lattice in Fig. 1(c)\nand the lattice in Fig. 1(b) reveal the NLM ferrimag-\nnetism in the ground state. Note also that the lattice\nshape at the edge under the open boundary condition\ndepicted in Fig. 1(c) is different from that in ref. VI [see\nFig. 1(b) in ref. VI]. Thus, one can recognize that the\nresults of the strip lattice with small width are irrespec-\ntive of boundary conditions. We also present clearly the\nexistence of the incommensurate modulation with long-\ndistance periodicity in the local magnetizations of both\nmodelsinFigs.1(b)and1(c). Ournumericalcalculations\nsuggest that the intermediate state of the 2D lattice in\nFig. 1(a) is the NLM ferrimagnetism.\nThis paper is organized as follows. In §2, we first\npresent our numerical calculation methods. In §3, we\nshow the ground-state properties of the lattice depicted\nin Fig. 1(c) in finite-size clusters. In §4, we show the\nground-state properties of the lattice depicted in Fig.\n1(b). Sections 5 and 6 are devoted to discussion and\nsummary, respectively.\nII. NUMERICAL METHODS\nWe employ two reliable numerical methods: the ex-\nact diagonalization (ED) method and the density matrixrenormalization group (DMRG) method32,33.\nThe ED method is used to obtain precise physical\nquantities of finite-size clusters. This method does not\nsuffer from the limitation of the cluster shape. It is ap-\nplicable even to systems with frustration, in contrast to\nthe quantum Monte Carlo (QMC) method coming across\nthe so-callednegative-signproblemforsystemswith frus-\ntration. The disadvantage of the ED method is the lim-\nitation that available system sizes are very small. Thus,\nwe should pay careful attention to finite-size effects in\nquantities obtained by this method.\nOntheotherhand,theDMRGmethodisverypowerful\nwhen a system is (quasi-)one-dimensionalunder the open\nboundary condition. The method can treat much larger\nsystemsthan theEDmethod. Notethatthe applicability\nof the DMRG method is irrespective of whether or not\nthe systems include frustrations. In the present research,\nwe use the “finite-system” DMRG method.\nIII. KAGOME STRIP LATTICE WITH SMALL\nWIDTH\nIn this section, we study the magnetic properties in\nthe ground state of the S= 1/2 Heisenberg model on the\nkagome strip lattice depicted in Fig. 1(c). The Hamilto-\nnian of this model is given by\nH=J1/summationdisplay\ni[Si,B·Si,C+Si,C·Si,A′+Si,C·Si+1,A+Si,C·Si+1,B]3\n+J2/summationdisplay\ni[Si,A·Si,B+Si,B·Si,A′]+JF/summationdisplay\ni[Si,A·Si+1,A+Si,A′·Si+1,A′], (1)\n\tB\n\tC\n0 10 20\nStotz–39.5–39–38.5–38E nergy kagome strip in Fig. 1(c) \nN=96, MS=900, SW=15\nJF/J 1=–1\nJ2/J 1=0.50J2/J 1=0.58\n0 0.5 1 1.5 2\nJ2/J 100.20.40.6M/M sN=24, periodic\nN=48, open\nN=96, open\nJF/J 1=–1kagome strip in Fig. 1(c) \nFIG. 2: (Color)(a) Dependence of the lowest energy on Sz\ntot.\nThe results of J2/J1= 0.5 and 0.58 for the system size of\nN= 96 are presented. Arrows indicate the values of the\nspontaneous magnetization Min eachJ2/J1. The position\nof an arrow is given by the highest Sz\ntotvalue among those\nthat give the common lowest energy. (b) J2/J1dependence\nofM/Msobtained from ED calculations for N= 24 (black\nsquare) under the periodic boundary condition and DMRG\ncalculation for N= 48 (red triangle) and 96 (blue inverted\ntriangle) under the open boundary condition. Note that for\nN= 96 in (a) and (b), we use MS= 900 and SW= 15 and,\nthat for N= 48 in (b), we use MS= 400 and SW= 15.\nwhereSi,ξistheS= 1/2spinoperatoratthe ξ-sublattice\nsite in the i-th unit cell. The positions of the four sub-\nlattices in a unit cell are denoted by A, A′, B, and C in\nFig. 1(c). Energies are measured in units of J1; we fixed\nJ1= 1 hereafter. In what follows, we examine the region\nof 0< J2/J1<∞in the case of JF=−1. Note that\nthe number of total spin sites is denoted by N; thus, the\nnumber of unit cells is N/4.\nWe examine the J2/J1dependence of the ratio M/Ms,\nwhereMandMsare the spontaneous and saturation\nmagnetizations, respectively. Let us explain the method\nusedtodetermine Masafunction of NandJ2/J1. First,\nwe calculate the lowest energy E(J2/J1,Sz\ntot,N), where\tB\n\tC\n0 10 20\ni–0.4 –0.2 00.20.4\nN=96 J F/J 1=–1 J 2/J 1=0.3 M=24\nAA’ BC\n0 10 20\ni–0.4 –0.2 00.20.4\nN=96 J F/J 1=–1 J 2/J 1=0.57 M=20 \nAA’ BC\nFIG. 3: (Color) Local magnetization /angbracketleftSz\ni,ξ/angbracketrightat each sublattice\nξ; A (black square), A′(red triangle), B (blue square), and C\n(purple inverted triangle). Panels (a) and (b) show results for\nJ2/J1= 0.3 and 0.57, respectively. These results are obtained\nfrom our DMRG calculations for N= 96 (i= 1,2,···,24).\nSz\ntotvalue is the z-component of the total spin. For ex-\nample, the energies for various Sz\ntotin the two cases of\nJ2/J1are presented in Fig. 2(a); the results are obtained\nby our DMRG calculations of the system of N= 96 with\nthe maximum number of retained states ( MS) of 600\nand the number of sweeps ( SW) of 15. The spontaneous\nmagnetization M(J2/J1,N) is determined as the highest\nSz\ntotamong those at the lowest common energy [see ar-\nrows in Fig. 2(a)]. Our results of the J2/J1dependence\nofM/Msare depicted in Fig. 2(b). We find the existence\nof the intermediate magnetic phase of 0 < M/M s<1/2\nbetween the LM ferrimagnetic phase of M/Ms= 1/2\nand the nonmagnetic phase. In order to determine the\nspin state of this intermediate phase, we calculate the\nlocal magnetization /angbracketleftSz\ni,ξ/angbracketright, where/angbracketleftA/angbracketrightdenotes the expec-\ntation value of the physical quantity AandSz\ni,ξis the\nz-component of Si,ξ. Figure 3 depicts our results for a\nsystem size N= 96 on the lattice depicted in Fig. 1(c)\nunder the open boundary condition; Figs. 3(a) and 3(b)\ncorrespondtothe caseofthe LMferrimagneticphaseand4\nthat of the intermediate phase, respectively.\nThe results clearly indicate the existence of the incom-\nmensurate modulation with long-distance periodicity in\nthe intermediate phase. We also confirm that the period-\nicities of the local magnetizations in the NLM ferrimag-\nnetism in the present model depend on the J2/J1value\nbut not on the length of the system, as reported in the\ncase of ref. VI.\nItisworthemphasizingthe pointthatthe intermediate\nphase commonly exists irrespective of the shape at the\nedge of the strip lattice when one compares the result\nof the present lattice depicted in Fig. 1(c) and that in\nref. VI. Therefore, wecan concludethat the intermediate\nphase exists and that it is NLM ferrimagnetic.\nAlthough there exists an intermediate NLM ferrimag-\nnetic phase in the case of the kagome strip lattice de-\npicted in Fig. 1(c), we should note that there is a large\ndiscrepancy in dimensionality between the kagome striplattice depicted in Fig. 1(c) and the 2D kagome lattice\ndepicted in Fig. 1(a). In the next section, we treat the\nkagome strip lattice depicted in Fig. 1(b) whose width is\nlarger than that of the kagome strip lattice depicted in\nFig. 1(c).\nIV. KAGOME STRIP LATTICE WITH LARGE\nWIDTH\nA. Hamiltonian\nIn this section, we study the ground-state properties\nof theS= 1/2 Heisenberg model on the kagome strip\nlattice depicted in Fig. 1(b). The Hamiltonian of this\nmodel is given by\nH=J1/summationdisplay\ni[Si,B·Si,C+Si,C·Si,D+Si,C·Si+1,A+Si,C·Si+1,B\n+Si,C′·Si,B′+Si,C′·Si,A′+Si,C′·Si+1,D+Si,C′·Si+1,B′]\n+J2/summationdisplay\ni[Si,A·Si,B+Si,B·Si,D+Si,D·Si,B′+Si,B′·Si,A′]\n+JF/summationdisplay\ni[Si,A·Si+1,A+Si,A′·Si+1,A′]\n−h/summationdisplay\ni[Sz\ni,A+Sz\ni,A′+Sz\ni,B+Sz\ni,B′+Sz\ni,C+Sz\ni,C′+Sz\ni,D]. (2)\nHere, the positions ofseven sublattices are denoted by A,\nA′, B, B′, C, C′, and D in Fig. 1(b). Note that the last\nterm of eq. 2 is the Zeeman term. The number of spin\nsitesisdenotedby N. Thenumberofunitcellsis N/7; we\nconsider N/14 as an integer. We mainly use the DMRG\nmethod for investigating the magnetic properties in the\nground state of this Q1D system under the open bound-\nary condition. We also investigate the properties under\nthe periodic boundary condition by the ED method, al-\nthough the size treated by this method is only in the case\nofN= 28. Hereafter, we consider J1= 1 as an energy\nscale and we investigate the region of 0 < J2/J1<∞in\nthe case of JF=−1.\nB. Phase diagram\nFirst, let us examine the J2/J1dependence of the ratio\nM/Msin the absence of the external magnetic field h.\nThe procedure for determining Mis the same as that\nmentioned in §3 [see also Fig. 4(a)]. We present our\nresults of the spontaneous magnetization in Fig. 4(b).\nWe successfully observe the intermediate-magnetizationregion irrespective of the boundary conditions. A careful\nobservation of Fig. 4(b) enables us to observe the eight\nregions at least in the finite-size system. As a matter\nof convenience, hereafter, we call these regions R 1, R2,\n···, R7, and R 8. In the case of N= 112 under the open\nboundarycondition, forexample, Fig. 5(a)illustratesthe\nregions R 1to R8: R1is the region of M/Ms= 3/7, R2is\nthe region of 11 /28≤M/Ms<3/7, R3is the region of\n1/8< M/M s<11/28, R4is the region of M/Ms= 1/8,\nR5is the region of 0 < M/M s<1/8, R6is the region of\nM/Ms= 0, R 7is the region of 0 < M/M s<1/7, and R 8\nis the region of M/Ms= 1/7. Here, the dashed lines in\nFig. 5(a) indicate the boundaries of these regions.\nIt should be noted that the values of M/Msin the\nR4region and that at the lower edge of the R 2region\nchange with increasing N, as shown in Fig. 4(b); the for-\nmer value is M/Ms= (N−14)/7Nand the latter value\nisM/Ms= (3N−28)/7N. These changes due to the in-\ncrease in system size come from the finite-size effect. We\nfind that the value of M/Msin the R 4region under the\nopen boundary condition increases and approaches the\nvalue of M/Ms= 1/7 whenNincreases. Furthermore,\nthe magnetization value in the R 4region is M/Ms= 1/75\n\tB\n \n\tC\n 0 10 20\nStotz–20020 E nergy kagome strip in Fig. 1(b) \nN=56, MS=600, SW=15 \nJF/J 1=–1\nJ2/J 1=0.20J2/J 1=0.57\n0 0.5 1 1.5 2\nJ2/J 100.20.40.6M/M skagome strip in Fig. 1(b) \nN=28, periodic \nN=56, open \nN=112, open\nJF/J 1=–1\nFIG. 4: (Color)(a) Lowest energy in each subspace divided by\nSz\ntot. The results of the DMRG calculations obtained when\nthe system size is N= 56 for J2/J1= 0.20 and 0.57 are\npresented. Arrows indicate the spontaneous magnetization M\nfor a given J2/J1. (b)J2/J1dependence of M/Msobtained\nfrom the ED calculations for N=28 (blue triangle) under the\nperiodic boundary condition and the DMRG calculations for\nN=56 (red square) and 112 (black pentagon) under the open\nboundary condition. Note that for N= 56 in (a) and (b),\nwe useMS= 600 and SW=15, and that for N= 112 in\n(b), we use MS≥900 and SW=15. Here, MSdenotes the\nmaximum number of retained states and SWthe number of\nsweeps used in DMRG calculations.\nin the case of N= 28 under the periodic boundary condi-\ntion. Therefore, it is expected that the value of M/Msin\nthe R4regionis 1 /7in the thermodynamic limit. We also\nconfirm that the value of M/Msat the lower edge of the\nR2region gradually increases and approaches the value\nofM/Ms= 3/7 with increasing N. In addition, we can-\nnot confirm the R 2region in the case of N= 28 under\nthe periodic boundary condition. These circumstances\nindicate a possibility that the R 2region merges with the\nR1region of M/Ms= 3/7 in the thermodynamic limit.\nNext, to determine whether each region survives in\nthe thermodynamic limit, we study the size depen-\ndences of the boundaries between the regions under the\nopen boundary condition. Figure 5(b) shows the re-\nsults of N=42, 56, 84, and 112 from DMRG calcu-\nlations. Note here that we define R 2as the region of(3N−28)/7N < M/M s<3/7 and R 4as the region of\nM/Ms= (N−14)/7Nin the finite-size system. One\ncan find immediately from Fig. 5(b) that all regions, ex-\ncept the R 7region, survive in the limit N→ ∞. To\ndetermine whether the R 7region survives in the thermo-\ndynamic limit, we investigate the size dependence of the\nwidth of the R 7region in Fig. 6. This plot shows us that\nthe width of the R 7region decreases with increasing N.\nIt is difficult to determine whether the R 7region survives\nin the thermodynamic limit. The convex downward be-\nhavior is observed for large sizes so that the region might\nsurvive; however, the observed behavior may be one of\ntheseriousfinite-sizeeffects. Theissueofestablishingthe\npresence or absence of the R 7region should be clarified\nin future studies.\n\tB\n\tC\n JF/J 1=–1\n0 0.01 0.02 0.03\n1/N00.511.52J2/J 1\nR1–R 2R2–R 3R3–R 4R4–R 5R5–R 6R6–R 7R7–R 80 0.5 1 1.5 2\nJ2/J 100.20.40.6M/M skagome strip in Fig. 1(b) \nN= 112, open \nR1 R2\nR3\nR4R5\nR6R7R8\nFIG. 5: (a) Definitions of the R 1-R8regions in the case of\nN= 112 underthe open boundary condition: R 1is the region\nofM/Ms= 3/7, R2is the region of 11 /28≤M/Ms<3/7,\nR3is the region of 1 /8< M/M s<11/28, R4is the region of\nM/Ms= 1/8, R5is the region of 0 < M/M s<1/8, R6is the\nregion of M/Ms= 0, R 7is the region of 0 < M/M s<1/7,\nand R 8is the region of M/Ms= 1/7. Dashed lines indi-\ncate the boundaries of these regions. (b) Size dependences\nof boundaries under the open boundary condition. The re-\nsults presented are those of N=42, 56, 84, and 112 from\nDMRG calculations. The curve line named R l-Rl+1indicates\nthe boundary line between the R land R l+1regions, where l\nis integer.6\n0 0.01 0.02 0.03 0.04\n1/N00.10.20.30.4Width of region R7 region\nFIG. 6: Size dependence of the width of the R 7region. It is\ndifficult to determine whether the R 7region survives in the\nthermodynamic limit.\nC. Magnetic properties in each region\nIn this subsection, we investigate the local magnetiza-\ntion/angbracketleftSz\ni,ξ/angbracketrightto study the magnetic structures in various\nregions except the R 2and R 6regions. Note that we cal-\nculate/angbracketleftSz\ni,ξ/angbracketrightwithin the subspace of the highest Sz\ntotcor-\nresponding to the spontaneous magnetization Mwhen\nJ2/J1is given. Considering the fact that the present\nlattice depicted in Fig. 1(b) has seven sublattices in the\nsystem, we will use different colors or symbols for each\nsublattice ξfor presenting our results of /angbracketleftSz\ni,ξ/angbracketright, as de-\npicted in the inset of Fig. 7(a); we use a black square for\nξ=A, a red triangle for ξ= A′, a blue cross for ξ= B, a\ngreen pentagon for ξ= B′, a purple inverted triangle for\nξ=C, an aqua diamond for ξ= C′, and a black circle for\nξ=D.\nFirst, we examine the R 1and R 8regions. We present\nour DMRG results of /angbracketleftSz\ni,ξ/angbracketrightof the system of N= 112 in\nFigs. 7(a) and 7(b) for J2/J1= 0.2 and 1.9, respectively.\nIn Fig. 7(a), we observethe uniform behavior of upward-\ndirection spins at the sublattice sites A, A′, B, B′, and\nD, and downward-direction spins at the sublattice sites\nC and C′. In Fig. 7(b), we also observe the uniform\nbehavior of upward-direction spins at the sublattice sites\nB, B′, C, and C′, and downward-direction spins at the\nsublatticesites A, A′, andD. Therefore, weconcludethat\nthe LM ferrimagnetic states are realized in the regions of\nR1and R 8.\nOur understanding of the origins of these LM ferri-\nmagnetic phases is based on the Marshall-Lieb-Mattis\n(MLM) theorem. In the case of J2/J1= 0, no frustra-\ntion occurs; thus, the spin state depicted in Fig. 8(a) is\nrealized. This state shows the LM ferrimagnetism with\nM/Ms= 3/7. The R 1region is directly connected to the\nLM ferrimagnetic state of J2/J1= 0. Therefore, the R 1\nregion of M/Ms= 3/7 is regarded as the LM ferrimag-\nnetic phase. In the limit J2→ ∞, on the other hand, the\npresent model becomes equal to a model of an S= 1/2\ndiamond chain depicted in Fig. 8(b). The value of mag-\tB\n\tC\n%\"\n# $\n#` $` \n\"` \n5 10 15\ni–0.2 00.20.4N=112 J F/J 1=–1 J 2/J 1=0.2 M=24\n5 10 15\ni–0.2 00.20.4N=112 J F/J 1=–1 J 2/J 1=1.9 M=8 \nFIG. 7: (Color) Local magnetization /angbracketleftSz\ni,ξ/angbracketrightat each sublattice\nξ. Panels (a) and (b) show results for J2/J1=0.2 and 1.9,\nrespectively. These results are obtained from our DMRG cal-\nculations for N= 112 (i=1, 2,···, 16). The inset in the\npanel (a) presents the correspondence relationship betwee n\neach colored symbol and each sublattice ξused in Figs. 7, 9,\nand 10.\nnetization takes M/Ms= 1/7 in the ground state of this\ndiamond chain according to the MLM theorem. There-\nfore, the R 8region of M/Ms= 1/7 is regarded as the LM\nferrimagnetic phase.\nNext, we investigate the R 3, R5, and R 7regions. Our\nresultsobtainedfromthe DMRG calculationsof N= 168\nare depicted in Figs. 9(a)-9(c) for J2/J1= 0.57, 1.14,\nand 1.69, respectively. We clearly observe incommensu-\nratemodulationswithlong-distanceperiodicitiesinevery\ncase in Fig. 9. In addition, we confirm from Fig. 4(b)\nthat the ratio M/Mschanges gradually with the varia-\ntion inJ2/J1in the R 3, R5, and R 7regions. Since the\nwidths of the R 3and R 5regions survive in the thermo-\ndynamic limit as was clarified in the previous subsection,\nwe conclude that the R 3and R 5regions are NLM ferri-\nmagnetic phases. Although it is unclear whether the R 7\nregion survives in the thermodynamic limit, this region\nis an NLM ferrimagnetic phase if it survives.\nFinally, in this subsection, we examine the R 4region.\nOur result of /angbracketleftSz\ni,ξ/angbracketrightforJ2/J1= 1 in the system of\nN= 168 is depicted in Fig. 10. We do not detect the in-\ncommensuratemodulation in this R 4region. In addition,\nwe confirm from Fig. 4(b) that the ratio M/Msin the\nR4region does not change with the variation in J2/J1in\ncontrast to the cases in the R 3and R 5regions. These7\n%\n\"` #` \n$` \"\n#\n$\n$\n$` \tB\n\tC\nFIG. 8: (a) Kagome strip lattice depicted in Fig. 1(b) in the\nlimit of J2/J1= 0. White arrowheads denote the classical\nspin configuration in the LM ferrimagnetic state of M/Ms=\n3/7. (b) Kagome strip lattice depicted in Fig. 1(b) in the\nlimit of J2→ ∞. White circles represent the effective S=\n1/2 spins formed by five S= 1/2 spins at the sublattices\nξ=A, A′, B, B′, and D. The black bold and black dotted lines\nshow the antiferromagnetic and ferromagnetic interaction s,\nrespectively.\ncircumstances suggest that the R 4region is the LM ferri-\nmagnetic phase. However, one cannot speculate the spin\nstate from the result of /angbracketleftSz\ni,ξ/angbracketrightbecause it is difficult to\nclarify the spin state on the basis of the results of deter-\nmining whether the spins are up or down, as was success-\nfully observed in the R 8region. We further investigate\nthe magnetization curve in this region to know whether\nthe R4region is the LM ferrimagnetic phase. The mag-\nnetization curves of J2/J1= 1 forN= 56 and N= 112\ncalculated by the DMRG method are presented in Fig.\n11(a)34. Figure 11(b) is obtained by zooming the region\nofhnearh= 0. One can observe the existence of the\nmagnetization plateaus at the height of the spontaneous\nmagnetization for both system sizes. We also confirm\nthat the difference in the width of the plateau between\nN= 112 and 56 is very small. These features indicate\nthat the spin gap exists in the R 4region in the thermo-\ndynamic limit. If the R 4region is the NLM ferrimagnetic\nphase, no spin gap is present in this region. (It was re-\nported in ref. VI that the NLM ferrimagnetism is gapless\nas a response to a uniform magnetic field.) Therefore,\nwe conclude that the R 4region is the LM ferrimagnetic\nphase.\nV. DISCUSSION\nWe discuss the relationships between the interval 0 <\nJ2/J1≤1 of the kagome strip lattice depicted in Fig.\n1(b) and those of the spatially anisotropic kagome lat-\tB\n\tC\n\tD\n10 20 \ni–0.2 00.20.4\nN=168 J F/J 1=–1 J 2/J 1=0.57 M=28 \n10 20 \ni–0.2 00.20.4\nN=168 J F/J 1=–1 J 2/J 1=1.14 M=7\n10 20 \ni–0.2 00.20.4N=168 J F/J 1=–1 J 2/J 1=1.69 M=2\nFIG. 9: (Color) Local magnetization /angbracketleftSz\ni,ξ/angbracketrightat each sublat-\nticeξ. The correspondence relationship between each colored\nsymbol and each sublattice ξis described in the inset in Fig.\n7(a). Panels (a), (b), and (c) show results for J2/J1=0.57,\n1.14, and 1.69, respectively. These results are obtained fr om\nour DMRG calculations for N= 168 (i=1, 2,···, 24).\ntice studied in ref. VI, while we consider the case of\nanother new lattice, with a larger but finite width along\nthe direction perpendicular to the bonds of interaction\nJFin Fig. 1(b), namely, the strip width.\nThe ratio M/Msof the R 1and R2regionsis commonly\nM/Ms= 3/7 in the thermodynamic limit of the lattice\ndepicted in Fig. 1(b). The difference of M/Ms= 3/7\nfromM/Ms= 1/3 in the case of the LM ferrimagnetic\nphase of the spatially anisotropic kagome lattice in ref.\nVIis attributedtothe finitenessofthestripwidth. Thus,\nthe ratio M/Msapproaches M/Ms= 1/3 when the strip8\n10 20 \ni–0.2 00.20.4\nN=168 J F/J 1=–1 J 2/J 1=1 M=11\nFIG. 10: (Color) Local magnetization /angbracketleftSz\ni,ξ/angbracketrightat each sublattice\nξ. This result is obtained from our DMRG calculations for\nJ2/J1= 1 in the system of N= 168 (i=1, 2,···, 24). The\ncorrespondence relationship between each colored symbol a nd\neach sublattice ξis described in the inset in Fig. 7(a).\nwidth increases; the R 1and R 2regions of the present\nmodel depicted in Fig. 1(b) correspond to the LM ferri-\nmagnetic phase of the spatially anisotropic kagome lat-\ntice studied in ref. VI.\nThe ratio M/Ms= 1/7 atJ2/J1= 1 in the present\nmodel depicted in Fig. 1(b) may be related to the fact\nthat the model includes seven sublattices, namely, the\nstrip width is finite. At least at J2/J1= 1, on the other\nhand, it is widely believed that the spontaneous mag-\nnetization disappears in the limit of infinite width17–27.\nAlthough the relationship should be clarified in the ex-\namination of the case of the lattice with an even larger\nstrip width, there are two possibilities of the behavior\nnear the case of J2/J1= 1. One is the case when the\nratioM/Msdecreases with increasing strip width and fi-\nnally vanishes, while the LM ferrimagnetic phase, such\nas R4, survives in systems with larger strip widths. In\nthis case, the R 3region corresponds to the NLM phase\nofthe two-dimensionalmodel onthe spatiallyanisotropic\nkagome lattice. In the other case, the LM ferrimagnetic\nphase, such as R 4, becomes narrower with increasing\nstrip width, while the ratio M/Msdoes not vanish; fi-\nnally, the R 3and R 5regions merge with each other. In\nthis case, the value of J2/J1at the boundary between\nthe R5and R 6regions decreases across J2/J1= 1. In\nany cases, it is important to note that from our finding\nof an intermediate phase in all the three cases in Fig. 1\nbetween the ferrimagnetic phase and the nonmagnetic\nphase, the phase is considered to exist irrespective of the\nstrip width.\nFinally, one should note that the intermediate phase\nbetween the Lieb-Mattis ferrimagnetic and nonmagnetic\nstates is observed in other cases. However, this phase\nis not always ferrimagnetic. One of the cases observed is\nthecaseofthethree-legladdersystemformingastriplat-\ntice obtained by cutting off from the spatially anisotropic\ntriangular lattice in ref. VI, in which the properties of\ntheintermediatephasewereunclear. Sakai’sunpublished\tB\n\tC\n0 1 2 3\nh00.51M/M skagome strip in Fig. 1(b) \nN=56 J F/J 1=–1 J 2/J 1=1 \nN=112 J F/J 1=–1 J 2/J 1=1 \n0 0.2 0.4\nh00.10.20.3M/M sN=56 \nN=112\nFIG. 11: Magnetization curve in the R 4region. Panel (a) is\nobtained from the DMRG calculations of J2/J1= 1 forN=\n56 (triangle) and N= 112 (square). Panel (b) is obtained by\nzooming the region near h= 0 in panel (a).\nstudy suggests that the intermediate phase is nematic38.\nCareful examinations are required to investigate such an\nintermediate phase if it is found.\nVI. SUMMARY\nWe have studied the ground-state properties of the\nS= 1/2 Heisenberg models on the kagome strip lattices\ndepicted in Figs. 1(b) and 1(c) by the ED and DMRG\nmethods. As a common phenomenon in the ground state\nofboth cases,wehaveconfirmedthe existenceofthe non-\nLieb-Mattis ferrimagnetism between the Lieb-Mattis fer-\nrimagnetic phase and the nonmagnetic phase. We have\nclearly found incommensurate modulations with long-\ndistance periodicity in the non-Lieb-Mattis ferrimagnetic\nstate. The occurrence of the non-Lieb-Mattis ferrimag-\nnetism irrespective of strip width strongly suggests that\nthe intermediate state found in the case of the spatially\nanisotropic kagome lattice in two dimensions is the non-\nLieb-Mattis ferrimagnetism.9\nAcknowledgments\nWe thank Prof. Toru Sakai for letting us know his\nunpublished results. This work was partly supported\nby Grants-in-Aid (Nos. 20340096, 23340109, 23540388,\nand 24540348) from the Ministry of Education, Culture,\nSports, Science and Technology of Japan. This work\nwas partly supported by a Grant-in-Aid (No. 22014012)\nfor Scientific Research and Priority Areas “Novel Statesof Matter Induced by Frustration” from the Ministry of\nEducation, Culture, Sports, Science and Technology of\nJapan. Diagonalization calculations in the present work\nwere carried out using TITPACK Version 2 coded by\nH. Nishimori. DMRG calculations were carried out us-\ning the ALPS DMRG application39. Some computations\nwere performed using the facilities of the Supercomputer\nCenter, Institute for Solid State Physics, University of\nTokyo.\n1E. Lieb and D. Mattis: J. Math. Phys. 3(1962) 749.\n2W. Marshall: Proc. Roy. Soc. A 232(1955) 48.\n3K.Takano, K.Kubo, andH.Sakamoto: J.Phys.: Condens.\nMatter8(1996) 6405.\n4K. Okamoto, T. Tonegawa, Y. Takahashi, and M.\nKaburagi: J. Phys.: Condens. Matter 11(1999) 10485.\n5T. Tonegawa, K. Okamoto, T. Hikihara, Y. Takahashi, and\nM. Kaburagi: J. Phys. Soc. Jpn. 69(2000) Suppl. A, 332.\n6T. Sakai and K. Okamoto: Phys. Rev. B 65(2002) 214403.\n7N. B. Ivanov and J. Richter: Phys. Rev. B 69(2004)\n214420.\n8S. Yoshikawa and S. Miyashita: Stastical Physics of Quan-\ntum Systems:novel orders and dynamics , J. Phys. Soc. Jpn.\n74(2005) Suppl., p. 71.\n9K. Hida: J. Phys.: Condens. Matter 19(2007) 145225.\n10K. Hida and K. Takano: Phys. Rev. B 78(2008) 064407.\n11R. R. Montenegro-Filho and M. D. Coutinho-Filho: Phys.\nRev. B78(2008) 014418.\n12T.ShimokawaandH.Nakano: J. Phys.Soc.Jpn. 80(2011)\n043703.\n13T.ShimokawaandH.Nakano: J. Phys.Soc.Jpn. 80(2011)\n125003.\n14N. B. Ivanov, J. Richter, and D. J. J. Farnell: Phys. Rev.\nB66(2002) 014421.\n15R. F. Bishop, P. H. Y. Li, D. J. J. Farnell, and C. E.\nCampbell: Phys. Rev. B 82(2010) 024416.\n16H. Nakano, T. Shimokawa, and T. Sakai: J. Phys. Soc.\nJpn.80(2011) 033709.\n17P. Lecheminant, B. Bernu, C. Lhuillier, L. Pierre, and\nP. Sindzingre: Phys. Rev. B 56(1997) 2521.\n18Ch. Waldtmann, H.-U. Everts, B. Bernu, C. Lhuil-\nlier, P. Sindzingre, P. Lecheminant, and L. Pierre:\nEur. Phys. J. B 2(1998) 501.\n19K. Hida: J. Phys. Soc. Jpn. 70(2001) 3673.\n20D. C. Cabra, M. D. Grynberg, P. C. W. Holdsworth, and\nP. Pujol: Phys. Rev. B 65(2002) 094418.\n21J. Schulenberg, A. Honecker, J. Schnack, J. Richter, and\nH.-J. Schmidt: Phys. Rev. Lett. 88(2002) 0167207.\n22A.Honecker, J.Schulenberg, andJ.Richter: J.Phys.:Con-\ndens. Matter 16(2004) S749.\n23O. Cepas, C. M. Fong, P. W. Leung, and C. Lhuillier:\nPhys. Rev. B 78(2008) 140405(R).24P. Sindzingre and C. Lhuillier: Europhys. Lett. 88(2009)\n27009.\n25H. Nakano and T. Sakai: J. Phys. Soc. Jpn. 79(2010)\n053707.\n26T. Sakai and H. Nakano: Phys. Rev. B 83(2011)\n100405(R).\n27H. Nakano and T. Sakai: J. Phys. Soc. Jpn. 80(2011)\n053704.\n28P. Azaria, C. Hooley, P. Lecheminant, C. Lhuillier, and A.\nM. Tsvelik: Phys. Rev. Lett. 81(1998) 1694.\n29Ch. Waldtmann, H. Kreutzmann, U. Schollwock, K.\nMaisinger, andH.-U. Everts: Phys. Rev.B 62(2000) 9472.\n30J. Schulenburg, A. Honecker, J. Schnack, J. Richter, and\nH.-J. Schmidt: Phys. Rev. Lett. 88(2002) 167207.\n31T. Shimokawa and H. Nakano: J. Phys.: Conf. Ser. 320\n(2011) 012007.\n32S. R. White: Phys. Rev. Lett. 69(1992) 2863.\n33S. R. White: Phys. Rev. B 48(1993) 10345.\n34One can find the macroscopic jump in this magnetization\ncurvenear the saturation fieldas well as in the case ofsome\nfrustrated systems30,35. It is known that the occurrence\nmechanism of the magnetization jump can be understood\nfrom the viewpoint of the localized-magnon picture (see\nrefs. VI, VI, and VI).\n35J. Schnack, H.-J. Schmidt, A. Honecker, J. Schulenburg,\nand J. Richter: J. Phys.: Conf. Ser. 51(2006) 43.\n36M. E. Zhitomirsky and H. Tsunetsugu: Phys. Rev. B 70\n(2004) 100403.\n37S. Abe, T. Sakai, K. Okamoto, and K. Tutsui: J. Phys.:\nConf. Ser. 320(2011) 012015.\n38T. Sakai: private communications.\n39A. F. Albuquerque, F. Alet, P. Corboz, P. Dayal, A.\nFeiguin, L. Gamper, E. Gull, S. Gurtler, A. Honecker, R.\nIgarashi, M. Korner, A. Kozhevnikov, A. Lauchli, S. R.\nManmana, M. Matsumoto, I. P. McCulloch, F. Michel, R.\nM. Noack, G. Pawlowski, L. Pollet, T. Pruschke, U. Scholl-\nwock, S. Todo, S. Trebst, M. Troyer, P. Werner, and S.\nWessel: J. Magn. Magn. Mater. 310(2007) 1187 (see also\nhttp://alps.comp-phys.org)." }, { "title": "1803.00235v1.Calculating_the_Magnetic_Anisotropy_of_Rare_Earth_Transition_Metal_Ferrimagnets.pdf", "content": "Calculating the Magnetic Anisotropy of\nRare-Earth|Transition-Metal Ferrimagnets\nChristopher E. Patrick,1,\u0003Santosh Kumar,1Geetha Balakrishnan,1Rachel\nS. Edwards,1Martin R. Lees,1Leon Petit,2and Julie B. Staunton1\n1Department of Physics, University of Warwick,\nCoventry CV4 7AL, United Kingdom\n2Daresbury Laboratory, Daresbury, Warrington WA4 4AD, United Kingdom\n(Dated: March 2, 2018)\nAbstract\nMagnetocrystalline anisotropy, the microscopic origin of permanent magnetism, is often ex-\nplained in terms of ferromagnets. However, the best performing permanent magnets based on rare\nearths and transition metals (RE-TM) are in fact ferrimagnets, consisting of a number of mag-\nnetic sublattices. Here we show how a naive calculation of the magnetocrystalline anisotropy of\nthe classic RE-TM ferrimagnet GdCo 5gives numbers which are too large at 0 K and exhibit the\nwrong temperature dependence. We solve this problem by introducing a \frst-principles approach\nto calculate temperature-dependent magnetization vs. \feld (FPMVB) curves, mirroring the exper-\niments actually used to determine the anisotropy. We pair our calculations with measurements on\na recently-grown single crystal of GdCo 5, and \fnd excellent agreement. The FPMVB approach\ndemonstrates a new level of sophistication in the use of \frst-principles calculations to understand\nRE-TM magnets.\n1arXiv:1803.00235v1 [cond-mat.mtrl-sci] 1 Mar 2018High-performance permanent magnets, as found in generators, sensors and actuators, are\ncharacterized by a large volume magnetization and a high coercivity [1]. The coercivity |\nwhich measures the resistance to demagnetization by external \felds | is upper-bounded\nby the material's magnetic anisotropy [2], which in qualitative terms describes a preference\nfor magnetization in particular directions. Magnetic anisotropy may be partitioned into\ntwo contributions: the shape anisotropy, determined by the macroscopic dimensions of the\nsample, and the magnetocrystalline anisotropy (MCA), which depends only on the mate-\nrial's crystal structure and chemical composition. Horseshoe magnets provide a practical\ndemonstration of shape anisotropy, but the MCA is less intuitive, arising from the relativistic\nquantum mechanical coupling of spin and orbital degrees of freedom [3].\nPermanent magnet technology was revolutionized with the discovery of the rare-earth/transition-\nmetal (RE-TM) magnet class, beginning with Sm-Co magnets in 1967 [4] (whose high-\ntemperature performance is still unmatched [5]), followed by the world-leading workhorse\nmagnets based on Nd-Fe-B [6, 7]. With the TM providing the large volume magnetization,\ncareful choice of RE yields MCA values which massively exceed the shape anisotropy con-\ntribution [8]. RE-TM magnets are now indispensable to everyday life, but their signi\fcant\neconomic and environmental cost has inspired a global research e\u000bort aimed at replacing\nthe critical materials required in their manufacture [9].\nIn order to perform a targeted search for new materials it is necessary to fully understand\nthe huge MCA of existing RE-TM magnets. An impressive body of theoretical work based\non crystal \feld theory has been built up over decades [10], where model parameters are\ndetermined from experiment (e.g. Ref. [11]) or electronic structure calculations [12{14]. An\nalternative and increasingly more common approach is to use these electronic structure\ncalculations, usually based on density-functional theory (DFT), to calculate the material's\nmagnetic properties directly without recourse to the crystal \feld picture [15{19].\nCalculating the MCA of RE-TM magnets presents a number of challenges to electronic\nstructure theory. The interaction of localized RE-4 felectrons with their itinerant TM\ncounterparts is poorly described within the most widely-used \frst-principles methodology,\nthe local spin-density approximation (LSDA) [12]. Indeed, the MCA is inextricably linked to\norbital magnetism whose contribution to the exchange-correlation energy is missing in spin-\nonly DFT [20, 21]. MCA energies are generally a few meV per formula unit, necessitating\na very high degree of numerical convergence [22]. Finally, the MCA depends strongly on\n2temperature, so a practical theory of RE-TM magnets must go beyond zero-temperature\nDFT and include thermal disorder [23].\nEven when these signi\fcant challenges have been overcome, there is a more fundamental\nproblem. Experiments access the MCA indirectly, measuring the change in magnetization of\na material when an external \feld is applied in di\u000berent directions. By contrast, calculations\nusually access the MCA directly by evaluating the change in energy when the material is\nmagnetized in di\u000berent directions, with no reference to an external \feld. These experimental\nand computational approaches arrive at the same MCA energy provided one is studying a\nferro magnet. However, the majority of RE-TM magnets (and many other technologically-\nimportant magnetic materials) are ferrimagnets, i.e. they are composed of sublattices with\nmagnetic moments of distinct magnitudes and orientations. Crucially the application of\nan external \feld may introduce canting between these sublattices, a\u000becting the measured\nmagnetization. Thus the standard theoretical approach of ignoring the external \feld is hard\nto reconcile with real experiments on ferrimagnets.\nIn this Letter, through a combination of calculations and experiments, we provide the\nhitherto missing link between electronic structure theory and practical measurements of the\nMCA. Speci\fcally, we show how to directly simulate experiments by calculating, from \frst\nprinciples (FP), how the measured magnetization ( M) varies as a function of \feld ( B) applied\nalong di\u000berent directions and at di\u000berent temperatures. We apply our \\FPMVB\" approach\nto the RE-TM ferro and ferrimagnets YCo 5and GdCo 5, which are isostructural to the\ntechnologically-important SmCo 5[24] and, in the case of GdCo 5, a source of controversy in\nthe literature [25{35]. Pairing FPMVB with new measurements of the MCA of GdCo 5allows\nus to resolve this controversy. More generally, FPMVB enables a new level of collaboration\nbetween theory and experiment in understanding the magnetic anisotropy of ferrimagnetic\nmaterials.\nThe electronic structure theory behind FPMVB treats magnetic disorder at a \fnite tem-\nperatureTwithin the disordered local moment (DLM) picture [36, 37]. The methodology\nallows the calculation of the magnetization of each sublattice i,M i(T) =Mi(T)^M i, and the\ntorque quantity @F(T)=@^M i, whereFis an approximation to the temperature dependent\nfree energy. @F(T)=@^M iaccounts for the anisotropy arising from the spin-orbit interaction,\nwhile the contribution from the classical magnetic dipole interaction is computed numer-\nically [38]. Many of the technical details of the DFT-DLM calculations [36, 39{43] were\n3FIG. 1. Data points and \fts of dF=d\u0012 calculated for GdCo 5(blue, empty symbols; Gd and Co\nmoments held antiparallel) and YCo 5(green, \flled symbols), at 0 and 300 K.\ndescribed in our recent study of the magnetization of the same compounds [44]; the exten-\nsions to calculate the torques are described in Ref. [37]. The Gd-4 felectrons are treated\nwith the local self-interaction correction [43], and we have also implemented the orbital\npolarization correction [20] following Refs. [45, 46] using reported Racah parameters [47].\nDetails are given as Supplemental Material (SM) [48].\nYCo 5and GdCo 5crystallize in the CaCu 5structure, consisting of alternating hexagonal\nRCo 2c/Co 3glayers [24]. Y is nonmagnetic, while in GdCo 5the large spin moment of Gd\n(originating mainly from its half-\flled 4 fshell) aligns antiferromagnetically with the Co\nmoments. We now consider a \\standard\" calculation of the MCA based on a rigid rotation\nof the magnetization. If the Gd and Co moments are held antiparallel, GdCo 5is e\u000bectively\na ferromagnet with reduced moment MCo\u0000MGd. Then, from the hexagonal symmetry we\nexpect the angular dependence of the free energy to follow \u00141sin2\u0012+\u00142sin4\u0012+O(sin6\u0012),\nwhere\u0012is the polar angle between the crystallographic caxis and the magnetization direc-\ntion. The constants \u00141;\u00142determine the change in free energy \u0001 F, calculated e.g. from the\nforce theorem [49] or the torque dF=d\u0012 [50].\nIn Fig. 1 we show dF=d\u0012 calculated for ferromagnetic YCo 5and GdCo 5at 0 and 300 K.\nFitting the data to the derivative of the textbook expression, sin 2 \u0012(\u00141+ 2\u00142sin2\u0012), \fnds\u00141\nand\u00142to be positive (easy caxis) with\u00141an order of magnitude larger than \u00142. Considering\nexperimentally measured anisotropy constants in the literature, for YCo 5our\u00141value of\n3.67 meV (all energies are per formula unit, f.u.) at 0 K compares favorably to the values of\n3.6 and 3.9 meV reported in Refs. [28] and [51]. At 300 K, our value of 2.19 meV exhibits a\nslightly faster decay with temperature compared to experiment (2.6 and 3.0 meV), which we\nattribute to our use of a classical spin hamiltonian in the DLM picture [36, 44]. However, for\nGdCo 5our calculated values of \u00141show very poor agreement with experiments [26, 29]. First,\n4at 0 K we \fnd \u00141to be larger than YCo 5(4.26 meV), while experimentally the anisotropy\nconstant is much smaller (1.5, 2.1 meV). Second, we \fnd \u00141decreases with temperature\n(2.39 meV at 300 K) while experimentally the anisotropy constant increases (2.7, 2.8 meV).\nTo understand these discrepancies we must ask how the anisotropy energies were actually\nmeasured. Torque magnetometry provides an accurate method of accessing the MCA [52],\nbut is technically challenging in RE-TM magnets, which require very high \felds to reach\nsaturation [53]. Singular point detection [54] and ferromagnetic resonance [55] has also\nbeen used to investigate the MCA of polycrystalline and thin-\flm samples. However, the\nmost commonly-used method for RE-TM magnets, employed in Refs. [26, 29], is based on\nthe seminal 1954 work by Sucksmith and Thompson [56] on the anisotropy of hexagonal\nferromagnets. This work provides a relation between the measured magnetization Mab\nand \feldBapplied in the hard plane in terms of \u00141,\u00142and the easy axis magnetization\nM0[48, 56]:\n(BM 0=2)=(Mab=M0)\u0011\u0011=\u00141+ 2\u00142(Mab=M0)2: (1)\nFurther introducing m= (Mab=M0), equation 1 shows that a plot of \u0011againstm2should\nyield a straight line with \u00141as the intercept. Even though this \\Sucksmith-Thompson\nmethod\" was derived for ferromagnets, the technical procedure of plotting \u0011againstm2can\nbe performed also for ferrimagnets like GdCo 5[26, 29]. In this case, the quantity extracted\nfrom the intercept is an e\u000bective anisotropy constant Ke\u000bso, unlike YCo 5, the anisotropy\nconstants reported in Refs. [26, 29] are distinct from the \u00141values extracted from Fig. 1. As\nrecognized at the time of the original experiments [27{30], the reduced value of Ke\u000bwith\nrespect to\u00141of YCo 5is a \fngerprint of canting between the Gd and Co sublattices.\nMaking contact with previous experiments thus requires we obtain Ke\u000b. To this end\nwe have developed a scheme of calculating \frst-principles hard-plane magnetization vs. \feld\n(FPMVB) curves, on which we perform the Sucksmith-Thompson analysis to directly mirror\nthe experiments. The central concept of FPMVB is that at equilibrium, the torques from\nthe exchange, spin-orbit and dipole interactions must balance those arising from the external\n\feld. Then,\nB=@F(T)\n@\u0012i1\nMicos\u0012i+P\njsin\u0012j@M j\n@\u0012i: (2)\nThe magnetization at a given B;T is determined by the angle set f\u0012Gd;\u0012Co1;\u0012Co2;:::gwhich\nsatis\fes equation 2 for every magnetic sublattice. The spin-orbit interaction breaks the\n5FIG. 2. Magnetization of GdCo 5vs. applied magnetic \feld shown on a standard plot (left panel)\nor after the Sucksmith-Thompson analysis (eq. 1, right panel). Crosses/circles are calculated with\nmethods (i)/(ii) discussed in the text, and the area between them shaded as a guide to the eye.\nNote the two methods are e\u000bectively indistinguishable in the left panel. The dashed/solid lines are\ncalculated from the model free energies F1andF2. The right panel also shows the geometry of the\nmagnetization and \feld with respect to the crystallographic c-axis (thick gray arrow).\nsymmetry of the Co 3gatoms such that altogether there are four independent angles to vary\nfor GdCo 5. The second term in the denominator of equation 2 re\rects that the magnetic\nmoments themselves might depend on \u0012i(magnetization anisotropy). We have tested (i)\nneglecting this contribution and (ii) modeling the dependence as Mi(\u0012i) =M0i(1\u0000pisin2\u0012i),\nwhereM0iandpiare parameterized from our calculations.\nFigure 2 shows FPMVB curves of GdCo 5calculated using equation 2 with methods (i)\nand (ii), (crosses and circles) which yield virtually identical values of Ke\u000b. TheMvs.B\ncurves in the left panel resemble those of a ferromagnet where, as the temperature increases,\nit becomes easier to rotate the moments away from the easy axis so that a given B\feld\ninduces a larger magnetization. However, plotting \u0011againstm2in the right panel tells a\nmore interesting story. The e\u000bective anisotropy constant Ke\u000b(y-axis intercept) at 0 K is\n1.53 meV, much smaller than \u00141of YCo 5. Furthermore Ke\u000bincreases with temperature,\nto 1.74 meV at 300 K. Therefore, in contrast to the standard calculations of Fig. 1, the\nFPMVB approach reproduces the experimental behavior of Refs. [26, 29].\n6Our FPMVB calculations provide a microscopic insight into the magnetization process.\nFor instance at 0 K and 9 T, we calculate that the cobalt moments rotate away from the\neasy axis by 6.1\u000e. By contrast the Gd moments have rotated by only 3.9\u000e, i.e. the ideal\n180\u000eGd-Co alignment has reduced by 2.2\u000e(the geometry is shown in Fig. 2). We also \fnd\ncanting between the di\u000berent Co sublattices, but not by more than 0.1\u000eat both 0 and 300 K\n(the calculated angles as a function of \feld are shown in the SM [48]). This Co-Co canting\nis small thanks to the Co-Co ferromagnetic exchange interaction, which remains strong over\na wide temperature range [44]. The temperature dependence of Ke\u000bcan be traced to the\nfact that the easy axis magnetization M0of GdCo 5initially increases with temperature [44].\nEven ifMabincreases with temperature at a given \feld, a faster increase in M0can lead to\nan overall hardening in Ke\u000b(equation 1).\nWe assign the canting in GdCo 5to a delicate competition between the exchange interac-\ntion favoring antiparallel Co/Gd moments, uniaxial anisotropy favoring c-axis (anti)alignment,\nand the external \feld trying to rotate all moments into the hard plane. We can quantify\nthese interactions by looking for a model parameterization of the free energy F. Crucially\nwe can train the model with an arbitrarily large set of \frst-principles calculations exploring\nsublattice orientations not accessible experimentally, and test its performance against the\ntorque calculations of equation 2. Neglecting the 0.1\u000ecanting within the cobalt sublattices\ngives two free angles, \u0012Gdand\u0012Co. Including Gd-Co exchange A, uniaxial Co anisotropy\nK1;Coand a dipolar contribution S(\u0012Gd;\u0012Co) [31, 48] leads naturally to a two-sublattice\nmodel [30],\nF1(\u0012Gd;\u0012Co) =\u0000Acos(\u0012Gd\u0000\u0012Co) +K1;Cosin2\u0012Co\n+S(\u0012Gd;\u0012Co): (3)\nThe training calculations showed additional angular dependences not captured by F1, so we\nalso investigated:\nF2(\u0012Gd;\u0012Co) =F1(\u0012Gd;\u0012Co) +K2;Cosin4\u0012Co\n+K1;Gdsin2\u0012Gd: (4)\nAs discussed below the training calculations showed no strong evidence of Gd-Co exchange\nanisotropy [31{34].\nThe dashed (solid) lines in Fig. 2 are the calculated Mvs.Bcurves obtained by mini-\nmizingF1(2)\u0000P\niM i\u0001B. The second term includes magnetization anisotropy on the cobalt\n7FIG. 3. Anisotropy constants Ke\u000bvs. temperature of YCo 5(green) and GdCo 5(blue). The left\npanel shows calculations using equation 2 at 0 and 300 K (stars), or using parameterized model\nexpressions F1(diamonds) and F2(circles), and from Ref. [57] (YCo 5, squares). For GdCo 5we\nalso show in red \u00141extracted from \\standard\" calculations where the Gd and Co moments were\nheld rigidly antiparallel (cf. Fig. 1). The experimental data in the right panel was measured by\nus for GdCo 5(crosses, with shaded background) or taken from Refs. [26], [29] and [58] (squares,\ndashed lines, circles) and Refs [28] and [51] (green diamonds and dashed lines, YCo 5).\nmoments [48, 57]. On the scale of the left panel both F1andF2give excellent \fts to the\ntorque calculations, especially up to moderate \felds. The plot of \u0011againstm2reveals some\ndi\u000berences with F2giving a marginally improved description of the data, but F1already\ncaptures the most important physics.\nWe also applied the FPMVB approach to YCo 5, using equation 2 and the model for F\nintroduced in Ref. [57]. Then, parameterizing the models [48] over the temperature range\n0{400 K, calculating Mvs.Bcurves and extracting Ke\u000busing the Sucksmith-Thompson\nplots gives the results shown in the left panel of Fig. 3. We also show \u00141of GdCo 5to\nemphasize the di\u000berence between FPMVB calculations and the \\standard\" ones of Fig. 1.\nComparing Ke\u000bto previously-published experimental measurements on GdCo 5raises\nsome issues. First, the three studies in the literature report anisotropy constants which di\u000ber\nby as much as 1 meV [26, 29, 58]. Indeed there was controversy over whether the observed\nresults were evidence of an anisotropic exchange interaction between Gd and Co [31, 32] or\n8an artefact of poor sample stoichiometry [33, 34]. Furthermore the only study performed\nabove room temperature [26] reports without comment some peculiar behavior where Ke\u000b\nof GdCo 5exceeds that of YCo 5at high temperature [28], despite conventional wisdom that\nthe half-\flled 4 fshell of Gd does not contribute to the anisotropy.\nOur calculations do in fact show an excess in the rigid-moment anisotropy of GdCo 5of\n16% at 0 K (Fig. 1) compared to YCo 5. The authors of Refs. [29, 31] \ftted their experimental\ndata with a much larger excess of 50%, while the high-\feld study of Ref. [33] found (11\n\u000615)%, with the authors of that work attributing the di\u000berence to an improved sample\nstoichiometry [34]. Our calculated excess at 0 K is formed from two major contributions:\nthe dipole interaction energy, which accounts for 0.31 meV/f.u., and K1;Gd(equation 4) which\nwe found to be 24% the size of K1;Co. The nonzero value of K1;Gdis due to the 5 delectrons,\nwhose presence is evident from the Gd magnetization (7.47 \u0016Bat 0 K). We did not \fnd a\nsigni\fcant contribution from anisotropic exchange, which we tested in two ways: \frst by\nattempting to \ft a term A(1\u0000p0sin2\u0012Co) cos(\u0012Gd\u0000\u0012Co) to our training set of calculations,\nand also by computing Curie temperatures with the (rigidly antiparallel) magnetization\ndirected either along the coraaxes. We found the magnitude of the anisotropy ( p0) to be\nsmaller than 0.5% and negative at 0 K, and to decrease in magnitude as the temperature\nis raised. Consistently the Curie temperature was found to be only 1 K higher for aaxis\nalignment, which we do not consider signi\fcant.\nHowever, our calculations do not predict the Ke\u000bvalue of GdCo 5to exceed YCo 5. Indeed,\nin Fig. 3\u00141of GdCo 5approaches that of YCo 5at high temperatures, which is signi\fcant\nbecause\u00141provides an upper bound for Ke\u000b[32]. To resolve this \fnal puzzle we performed\nour own measurements of Ke\u000bon the single crystal whose growth we reported recently [44].\nHard and easy axis magnetization curves up to 7 T were measured in a Quantum Design\nsuperconducting quantum interference device (SQUID) magnetometer, and the anisotropy\nconstants extracted from Sucksmith-Thompson plots [48]. The right panel of Fig. 3 shows\nour newly measured data as crosses. Previously reported measurements are shown in faint\nblue/green for GdCo 5[26, 29, 58]/YCo 5[28, 51].\nUp to 200 K, there is close agreement between the experiments of Ref. [26], our own ex-\nperiments, and the FPMVB calculations. Above this temperature our new experiments show\nthe expected drop in Ke\u000b, while the previously reported data show a continued rise [26]. We\nrepeated our measurements using di\u000berent protocols and found a reasonably large variation\n9in the extracted Ke\u000b[48]. Even taking this variation into account as the shaded area in\nFig. 3, the drop is still observed.\nWe therefore do not believe the high temperature behavior reported in Ref. [26] has\nan intrinsic origin. Possible extrinsic factors include the method of sample preparation,\ndegradation of the RCo 5phase at elevated temperatures [59], and potential systematic error\nwhen extracting Ke\u000b. We note that even the idealized theoretical curves in Fig. 2 show\ncurvature at higher temperature, making it more di\u000ecult to \fnd the intercept.\nIn conclusion, we have introduced the FPMVB approach to interpret experiments mea-\nsuring anisotropy of ferrimagnets, particularly RE-TM permanent magnets. We presented\nthe method in the context of our DLM formalism, but any electronic structure theory ca-\npable of calculating magnetic couplings relativistically [60{64] should be able to produce\nFPMVB curves, at least at zero temperature. However standard calculations which neglect\nthe external \feld should be used with care when comparing to experiments on ferrimagnets.\nSimilarly, the prototype GdCo 5serves as a reminder that a simple view of the anisotropy\nenergy does not fully describe the magnetization processes in ferrimagnets, which might have\nimplications in understanding e.g. magnetization reversal in nano-magnetic assemblies [65].\nOverall our work demonstrates the bene\ft of interconnected computational and experimen-\ntal research in this key area.\nThe present work forms part of the PRETAMAG project, funded by the UK Engineer-\ning and Physical Sciences Research Council (EPSRC), Grant no. EP/M028941/1. Crystal\ngrowth work at Warwick is also supported by EPSRC Grant no. EP/M028771/1. Work at\nDaresbury Laboratory was supported by an EPSRC service level agreement with the Sci-\nenti\fc Computing Department of STFC. We thank E. Mendive-Tapia for useful discussions\nand A. Vasylenko for continued assistance in translating references.\n\u0003c.patrick.1@warwick.ac.uk\n[1] S. Chikazumi, Physics of Ferromagnetism , 2nd ed. (Oxford University Press, 1997).\n[2] H. Kronm uller, Phys. Stat. Sol. b 144, 385 (1987).\n[3] P. Strange, Relativistic Quantum Mechanics (Cambridge University Press, 1998).\n[4] K. Strnat, G. Ho\u000ber, J. Olson, W. Ostertag, and J. J. Becker, J. Appl. Phys. 38, 1001 (1967).\n10[5] O. Gut\reisch, M. A. Willard, E. Br uck, C. H. Chen, S. G. Sankar, and J. P. Liu, Adv. Mater.\n23, 821 (2011).\n[6] M. Sagawa, S. Fujimura, N. Togawa, H. Yamamoto, and Y. Matsuura, J. Appl. Phys. 55,\n2083 (1984).\n[7] J. J. Croat, J. F. Herbst, R. W. Lee, and F. E. Pinkerton, J. Appl. Phys. 55, 2078 (1984).\n[8] J. M. D. Coey, IEEE Trans. Magn. 47, 4671 (2011).\n[9] R. Skomski, P. Manchanda, P. Kumar, B. Balamurugan, A. Kashyap, and D. J. Sellmyer,\nIEEE Trans. Magn. 49, 3215 (2013).\n[10] M. D. Kuz'min and A. M. Tishin, in Handbook of Magnetic Materials , Vol. 17, edited by\nK. H. J. Buschow (Elsevier B.V., 2008) Chap. 3, p. 149.\n[11] Z. Tie-song, J. Han-min, G. Guang-hua, H. Xiu-feng, and C. Hong, Phys. Rev. B 43, 8593\n(1991).\n[12] M. Richter, J. Phys. D: Appl. Phys. 31, 1017 (1998).\n[13] M. D. Kuz'min, Y. Skourski, D. Eckert, M. Richter, K.-H. M uller, K. P. Skokov, and I. S.\nTereshina, Phys. Rev. B 70, 172412 (2004).\n[14] P. Delange, S. Biermann, T. Miyake, and L. Pourovskii, Phys. Rev. B 96, 155132 (2017).\n[15] L. Steinbeck, M. Richter, and H. Eschrig, J. Magn. Magn. Mater. 226{230, Part 1 , 1011\n(2001).\n[16] P. Larson, I. I. Mazin, and D. A. Papaconstantopoulos, Phys. Rev. B 69, 134408 (2004).\n[17] H. Pang, L. Qiao, and F. S. Li, Phys. Status Solidi B 246, 1345 (2009).\n[18] M. Matsumoto, R. Banerjee, and J. B. Staunton, Phys. Rev. B 90, 054421 (2014).\n[19] A. Landa, P. S oderlind, D. Parker, D. \u0017Aberg, V. Lordi, A. Perron, P. E. A. Turchi, R. K.\nChouhan, D. Paudyal, and T. A. Lograsso, \\Thermodynamics of the SmCo 5compound doped\nwith Fe and Ni: an ab initio study,\" (2017), arXiv:1707.09447.\n[20] O. Eriksson, B. Johansson, R. C. Albers, A. M. Boring, and M. S. S. Brooks, Phys. Rev. B\n42, 2707 (1990).\n[21] H. Eschrig, M. Sargolzaei, K. Koepernik, and M. Richter, Europhys. Lett. 72, 611 (2005).\n[22] G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuurmans, Phys. Rev. B 53, 14415 (1996).\n[23] J. B. Staunton, S. Ostanin, S. S. A. Razee, B. L. Gyor\u000by, L. Szunyogh, B. Ginatempo, and\nE. Bruno, Phys. Rev. Lett. 93, 257204 (2004).\n[24] K. Kumar, J. Appl. Phys. 63, R13 (1988).\n11[25] K. Buschow, A. van Diepen, and H. de Wijn, Solid State Commun. 15, 903 (1974).\n[26] A. Ermolenko, IEEE Trans. Mag. 12, 992 (1976).\n[27] S. Rinaldi and L. Pareti, J. Appl. Phys. 50, 7719 (1979).\n[28] A. S. Yermolenko, Fiz. Metal. Metalloved. 50, 741 (1980).\n[29] R. Ballou, J. D\u0013 eportes, B. Gorges, R. Lemaire, and J. Ousset, J. Magn. Magn. Mater. 54,\n465 (1986).\n[30] R. Radwa\u0013 nski, Physica B+C 142, 57 (1986).\n[31] R. Ballou, J. D\u0013 eportes, and J. Lemaire, J. Magn. Magn. Mater. 70, 306 (1987).\n[32] P. Gerard and R. Ballou, J. Magn. Magn. Mater. 104, 1463 (1992).\n[33] R. Radwa\u0013 nski, J. Franse, P. Quang, and F. Kayzel, J. Magn. Magn. Mater. 104, 1321 (1992).\n[34] J. J. M. Franse and R. J. Radwa\u0013 nski, in Handbook of Magnetic Materials , Vol. 7, edited by\nK. H. J. Buschow (Elsevier North-Holland, New York, 1993) Chap. 5, p. 307.\n[35] T. Zhao, H. Jin, R. Gr ossinger, X. Kou, and H. R. Kirchmayr, J. Appl. Phys. 70, 6134 (1991).\n[36] B. L. Gy or\u000by, A. J. Pindor, J. Staunton, G. M. Stocks, and H. Winter, J. Phys. F: Met.\nPhys. 15, 1337 (1985).\n[37] J. B. Staunton, L. Szunyogh, A. Buruzs, B. L. Gyor\u000by, S. Ostanin, and L. Udvardi, Phys.\nRev. B 74, 144411 (2006).\n[38] We performed a sum over dipoles [1] using the calculated Mi(T) out to a radius of 20 nm.\n[39] E. Bruno and B. Ginatempo, Phys. Rev. B 55, 12946 (1997).\n[40] P. Strange, J. Staunton, and B. L. Gyor\u000by, J. Phys. C: Solid State Phys. 17, 3355 (1984).\n[41] M. D ane, M. L uders, A. Ernst, D. K odderitzsch, W. M. Temmerman, Z. Szotek, and W. Herg-\nert, J. Phys.: Condens. Matter 21, 045604 (2009).\n[42] S. H. Vosko, L. Wilk, and M. Nusair, Can. J. Phys. 58, 1200 (1980).\n[43] M. L uders, A. Ernst, M. D ane, Z. Szotek, A. Svane, D. K odderitzsch, W. Hergert, B. L.\nGy or\u000by, and W. M. Temmerman, Phys. Rev. B 71, 205109 (2005).\n[44] C. E. Patrick, S. Kumar, G. Balakrishnan, R. S. Edwards, M. R. Lees, E. Mendive-Tapia,\nL. Petit, and J. B. Staunton, Phys. Rev. Materials 1, 024411 (2017).\n[45] H. Ebert and M. Battocletti, Solid State Commun. 98, 785 (1996).\n[46] H. Ebert, \\Fully relativistic band structure calculations for magnetic solids - formalism and\napplication,\" in Electronic Structure and Physical Properies of Solids: The Uses of the LMTO\nMethod Lectures of a Workshop Held at Mont Saint Odile, France, October 2{5,1998 , edited\n12by H. Dreyss\u0013 e (Springer Berlin Heidelberg, Berlin, Heidelberg, 2000) pp. 191{246.\n[47] L. Steinbeck, M. Richter, and H. Eschrig, Phys. Rev. B 63, 184431 (2001).\n[48] See Supplemental Material for further experimental and computational details, description\nof the orbital polarization correction, discussion of magnetization anisotropy, the parameters\nused to simulate MvBcurves and the angle sets which satisfy equation 2.\n[49] G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuurmans, Phys. Rev. B 41, 11919 (1990).\n[50] X. Wang, R. Wu, D.-s. Wang, and A. J. Freeman, Phys. Rev. B 54, 61 (1996).\n[51] J. M. Alameda, D. Givord, R. Lemaire, and Q. Lu, J. Appl. Phys. 52, 2079 (1981).\n[52] H. Klein, A. Menth, and R. Perkins, Physica B+C 80, 153 (1975).\n[53] K. H. J. Buschow and F. R. de Boer, Physics of Magnetism and Magnetic Materials (Springer,\nBoston, MA, 2003).\n[54] A. Paoluzi, L. Pareti, M. Solzi, and F. Albertini, J. Magn. Magn. Mater. 132, 185 (1994).\n[55] Y. Wang, Y. Zhao, S. Wang, M. Lu, H. Zhai, Y. Zhai, K. Shono, and X. Yu, J. Appl. Phys.\n93, 7789 (2003).\n[56] W. Sucksmith and J. E. Thompson, Proc. Royal Soc. A 225, 362 (1954).\n[57] J. Alameda, J. D\u0013 eportes, D. Givord, R. Lemaire, and Q. Lu, J. Magn. Magn. Mater. 15, 1257\n(1980).\n[58] T. Katayama, M. Ohkoshi, Y. Koizumi, T. Shibata, and T. Tsushima, Appl. Phys. Lett. 28,\n635 (1976).\n[59] F. Den Broeder and K. Buschow, J. Less-Common Metals 29, 65 (1972).\n[60] L. Udvardi, L. Szunyogh, K. Palot\u0013 as, and P. Weinberger, Phys. Rev. B 68, 104436 (2003).\n[61] H. Ebert and S. Mankovsky, Phys. Rev. B 79, 045209 (2009).\n[62] J. Hu and R. Wu, Phys. Rev. Lett. 110, 097202 (2013).\n[63] S. Ayaz Khan, P. Blaha, H. Ebert, J. Min\u0013 ar, and O. \u0014Sipr, Phys. Rev. B 94, 144436 (2016).\n[64] M. Ho\u000bmann, B. Zimmermann, G. P. M uller, D. Sch urho\u000b, N. S. Kiselev, C. Melcher, and\nS. Bl ugel, Nat. Commun. 8, 308 (2017).\n[65] Z. J. Guo, J. S. Jiang, J. E. Pearson, S. D. Bader, and J. P. Liu, Appl. Phys. Lett. 81, 2029\n(2002).\n13" }, { "title": "1905.03521v1.Bidirectional_spin_wave_driven_domain_wall_motion_in_antiferromagnetically_coupled_ferrimagnets.pdf", "content": "1 \n Bidirectional spin-wave -driven domain wall motion in \nantiferromagnetically coupled ferrimagnets \nSe-Hyeok Oh1, Se Kwon Kim2, Jiang Xiao3,4,5, and Kyung- Jin Lee1,6,7 * \n1Department of Nano -Semiconductor and Engineering, Korea University, Seoul 02841, Korea \n2Department of Physics and Astronomy, University of Missouri, Columbia , Missouri 65211 , USA \n3Department of Physics and State Key Laboratory of Surface Physics, Fudan University, Shang hai \n200433, China \n4Collaborative Innovation Center of Advanced Microstructures, Nanjing 210093, China \n5Institute for Nanoelectronics Devices and Quantum Computing, Fudan University, Shanghai 200433, \nChina \n6Department of Materials Science and Engineering, K orea University, Seoul 02841, Korea \n7KU-KIST Graduate School of Converging Science and Technology, Korea University, Seoul 02841, \nKorea \n \n* corresponding email: kj_lee@korea.ac.kr \n 2 \n Abstract \nWe investigate ferrimagnetic domain wall dynamics induced by circular ly polarized \nspin wave s theoretically and numerically . We find that the direction of domain wall \nmotion depends on both the circular polarization of spin wave s and the sign of net spin \ndensity of ferrimagnet. Below the angular momentum compensation point, left - (right-) \ncircularly polarized spin wave s push a domain wall toward s (away from ) the spin -wave \nsource. Above the angular momentum compensation point, on th e other hand, the \ndirection of domain wall motion is reversed . This bidirectional motion originate s from \nthe fact that the sign of spin -wave -induced magnon ic torque depends on the circular \npolarization and the subsequent response of the domain wall to the magnonic torque is \ngoverned by the net spin density . Our finding provides a way to utilize a spin wave as a \nversatile driving force for bidirectional domain wall motion . \n \nI. Introduction \nOne class of ferrimagnet s of emerging interest is a rare- earth (RE) -transition metal (TM) \ncompound where the RE and TM moments are coupled antiferromagnetically. Owing to \ndifferent Land é-g factors between the RE and TM elements, RE-TM ferr imagnet s exhibit two \nunique compensation temperatures : the magnetic moment compensation temperature 𝑇𝑇𝑀𝑀 at \nwhich net magnetic moment vanishes , and the angular momentum compensation temperature \n𝑇𝑇𝐴𝐴 at which net angular momentum vanishes [1-3]. \nResearch on ferr imagnetic materials ha s focused on the understanding of their \nfundamental magnetism [4] and optical switching of magnetization [5 -11]. Recently, RE -TM \nferrimagnet s attract renewed interest as they offer a material platform to investigate the \nantiferromagnetic spintronics [ 12-15]. Compared to ferromagnets that have served as core \nmaterial s for spintronics research, antiferromagnets exhibit several distinct features such as \nthe immunity to external field perturbations and fast dynamics due to antiferromagnetic \nexchange interaction. However, the external -field immunity of true antiferromagnets results \nin the experimental difficulty in both creating and controlling antiferromagnetic textures. On \nthe other hand, RE -TM ferrimagnets have finite magnetic moment at the angular momentum 3 \n compensation point at which the antiferromagnetic dynamics is realized. As a result, \npreviously established creation and detection schemes for ferromagnets is directly appli cable \nto RE -TM ferrimagnets. This simple but strong benefit of RE -TM ferrimagnets have recently \ninitiated extensive studies on ferrimagnets, which include magnetization switching [ 16-19], \ndomain wall motion [ 20-24], skyrmion (or bubble domain) motion [ 25-28], low damping [ 29], \nand efficient spin -transfer and spin- orbit torque s due to antiferromagnetic alignment of \natomic spins [30,31 ]. \nAmong the previous studies listed above, the low damping of RE -TM ferrimagnets [2 9] \nis of particular interest from the view point of magnonic applications based on ferrimagnets \nbecause it enables a long -distance propagation of spin waves (SWs) . For ferromagnets [ 32-39] \nand antiferromagnets [ 40-43], it was reported that a SW can move a DW by transferring its \nangular momentum or linear momentum . Though the SW property in ferrimagnet s was \nestablished [44-47], the effect of SWs on ferrimagnetic domain wall (DW) motion remains \nunexplored. In comparison to ferromagnets and antiferromagnets, antiferromagnetically \ncoupled ferrimagnets exhibit a distinguishing feature of SW eigenmode s. In ferromagnet s, a \nspin wave (SW) with only one type of polarization is permitted , which drives a DW toward s \nthe SW source through the angular momentum trans fer [3 4-37]. In antiferromagnet s, however, \nboth the left - and right -circular ly polarized SWs are allowed and energetically degenerate, \nwhich can transfer the linear momentum to a DW through the SW reflection [40,41,43] , \nresulting in the DW motion away from the SW source . In antiferromagnetically coupled \nferrimagnet s, on the other hand, the degeneracy of the two circular ly polarized SWs can be \nlifted depending on the net spin density of ferrimagnet . Given that SW -induced DW motion \nin ferrimagnets has been unexplored, interesting and important question s remain unanswered : \nhow a SW m oves a ferrimagnetic DW and what the role of circular polarization of SW is . \nIn this paper, we study the dynamics of a ferrimagnetic DW induced by a SW in the \nvicinity of the ang ular momentum compensation temperature 𝑇𝑇𝐴𝐴. We inve stigate DW \ndynamics induced by left - and right -circular ly polarized SWs [see Fig. 1(a) for an illustration \nof the two eigenmodes ]. We begin with theoretical analysis based on the Lagrangian density \nand SW dispersion. We then conduct numerical simulation based on the atomistic Landau -\nLifshitz -Gilbert (LLG) equation to confirm the analytical results . Our model system is shown 4 \n in Fig. 1(b). \n \nII. Model \nOur model system is a simple bipartite ferrimagnet which consists of two sublattices \nlabeled by A and B. We introduce the staggered vector 𝒏𝒏=(𝑨𝑨k−𝑩𝑩k)/2, and 𝒎𝒎=𝑨𝑨k+\n𝑩𝑩k, where 𝑨𝑨k and 𝑩𝑩k are the unit vectors of spin moment at a site k that belongs to the \nsublattice s A and B, respectively. The Lagrangian density for the ferrimagnet is given by \n[26,47- 49] \nℒ=[−s𝒏𝒏̇∙(𝒏𝒏×𝒎𝒎)−𝛿𝛿𝑠𝑠𝐚𝐚(𝒏𝒏)∙𝒏𝒏̇]−𝒰𝒰, (1) \nwhere 𝑠𝑠=(𝑠𝑠𝐴𝐴+𝑠𝑠𝐵𝐵)/2, 𝛿𝛿𝑠𝑠=𝑠𝑠𝐴𝐴−𝑠𝑠𝐵𝐵, 𝑠𝑠𝑖𝑖=𝑀𝑀𝑖𝑖/𝛾𝛾𝑖𝑖 is the angular momentum density , 𝑀𝑀𝑖𝑖 \nis the magnetic moment, 𝛾𝛾𝑖𝑖 is the gyromagnetic ratio for sublattice 𝑖𝑖, 𝐚𝐚(𝒏𝒏) is the vector \npotential for the magnetic monopole . The total energy 𝒰𝒰 includes the exchange energy and \nanisotropy energy as \n𝒰𝒰=𝑎𝑎\n2|𝒎𝒎|2+𝐴𝐴\n2(∇𝒏𝒏)2−𝐾𝐾\n2(𝒛𝒛�∙𝒏𝒏)2+𝜅𝜅\n2(𝒙𝒙�∙𝒏𝒏)2, (2) \nwhere 𝑎𝑎 is the homogeneous exchange, 𝐴𝐴 is the inhomogeneous exchange, 𝐾𝐾 is the easy-\naxis anisotropy constant , and 𝜅𝜅 is the hard-axis anisotropy constant. The Rayleigh function \naccounting for the dissipation is given by ℛ=𝛼𝛼𝑠𝑠𝒏𝒏̇2 where 𝛼𝛼 is the Gilbert damping \nconstant. From the Lagrangian density and the Rayleigh dissipation , we obtain the equation \nof motion in terms of staggered vector 𝒏𝒏 by integrating out the net magnetization variable \n𝒎𝒎 [26]: \n𝜌𝜌𝒏𝒏×𝒏𝒏̈+2𝛼𝛼𝑠𝑠𝒏𝒏×𝒏𝒏̇+𝛿𝛿𝑠𝑠𝒏𝒏̇=𝒏𝒏×𝒇𝒇𝒏𝒏, (3) \nwhere 𝜌𝜌=𝑠𝑠2/𝑎𝑎 paramet rizes the inertia and 𝒇𝒇𝒏𝒏=−𝛿𝛿𝒰𝒰𝛿𝛿𝒏𝒏⁄ is the effective field. After \nlinearizing the equations for small- amplitude fluctuations from the uniform state, w e consider \nthe SW ansatz as 𝒏𝒏 (𝑥𝑥,𝑡𝑡)=Re��𝑛𝑛𝑥𝑥exp[𝑖𝑖(𝜔𝜔𝑡𝑡−𝑘𝑘𝑥𝑥)],𝑛𝑛𝑦𝑦exp[𝑖𝑖(𝜔𝜔𝑡𝑡−𝑘𝑘𝑥𝑥)],1��, where \n𝑛𝑛𝑥𝑥,𝑛𝑛𝑦𝑦 are the amplitude s of SW (|𝑛𝑛𝑥𝑥|,�𝑛𝑛𝑦𝑦�≪1), 𝜔𝜔 is the SW frequency, and 𝑘𝑘 is the \nwavevector. By solving the linearized equations with this ansatz, we obtain the d ispersion 5 \n relation as \n𝜔𝜔±=±𝛿𝛿𝑠𝑠+�𝛿𝛿𝑠𝑠2+4𝜌𝜌(𝐴𝐴𝑘𝑘2+𝐾𝐾+𝜅𝜅/2)\n2𝜌𝜌. (4). \nHere the u pper (lower) sign corresponds to the left- (right -) circular ly polarized SW. The \nresonance frequencies for left - and right -circular ly polarized SWs are different except at the \nangular mom entum compensation point 𝑇𝑇𝐴𝐴 where the net spin density 𝛿𝛿𝑠𝑠 is zero . Figure 2 \nshows the agreement between the analytic SW dispersion relations [Eq. (4) ; lines ] and \nnumerical results that will be discussed below (symbols) . Below or above 𝑇𝑇𝐴𝐴, the energy of \nright -circular ly polarized SW differs from that of left- circular ly polarized SW [see Fig. 2(a) \nand Fig. 2(c) ]. At 𝑇𝑇𝐴𝐴 [Fig. 2(b) ], two circular ly polarized SW s are degenerate, which is \nanalogous with antiferromagnetic SW s. \nWe n ext look into the dynamics of ferrimagnetic DW induced by SW s. We consider \n𝒏𝒏 as 𝒏𝒏=𝒏𝒏𝟎𝟎+𝜹𝜹𝒏𝒏 with DW texture 𝒏𝒏𝟎𝟎 and small fluctuation 𝜹𝜹𝒏𝒏 (|𝜹𝜹𝒏𝒏|≪|𝒏𝒏𝟎𝟎|) with the \nconstraint 𝒏𝒏𝟎𝟎∙𝜹𝜹𝒏𝒏=0 to keep the unit length of 𝒏𝒏 to linear order in 𝜹𝜹𝒏𝒏. We introduc e two \ncollective coordinates [50], the DW position 𝑋𝑋(𝑡𝑡) and center angle 𝜙𝜙(𝑡𝑡), and define a DW \n𝒏𝒏𝟎𝟎 by Walker ansatz [51], 𝒏𝒏𝟎𝟎(𝑥𝑥,𝑡𝑡)=(sin𝜃𝜃sin𝜙𝜙,sin𝜃𝜃cos𝜙𝜙,𝑐𝑐𝑐𝑐𝑠𝑠𝜃𝜃 ) where 𝜃𝜃=\n2tan−1[exp{(𝑥𝑥−𝑋𝑋)/𝜆𝜆}] and 𝜆𝜆 is the DW width . We consider the magnonic torque 𝝉𝝉𝒎𝒎 \nthat is given by [34,37,39,52] \n𝝉𝝉𝒎𝒎=−𝐴𝐴[(𝑱𝑱𝒎𝒎∙𝜵𝜵)𝒏𝒏𝟎𝟎−(𝜕𝜕𝑥𝑥𝜌𝜌𝑚𝑚)𝒏𝒏𝟎𝟎×𝜕𝜕𝑥𝑥𝒏𝒏𝟎𝟎], (5) \nwhere magnon -flux density 𝐽𝐽𝑚𝑚𝑥𝑥=𝒏𝒏𝟎𝟎∙〈𝜹𝜹𝒏𝒏×𝜕𝜕𝑥𝑥𝜹𝜹𝒏𝒏〉, and magnon number density 𝜌𝜌𝑚𝑚=\n〈𝜹𝜹𝒏𝒏〉2/2. The first term in Eq. (5) represents the adiabatic magnonic torque rooted in the \nmagnon current , and the second term represents the non-adiabatic magnonic torque caused by \nthe gradient of the magnon density . Inserting Eq. (5) into the staggered LLG equation Eq. (3), \nwe derive two coupled equations of motion as \n𝑀𝑀𝑋𝑋̈−𝐺𝐺𝜙𝜙̇+𝑀𝑀𝑋𝑋̇𝜏𝜏⁄=𝐹𝐹𝑚𝑚, (6) \n𝐼𝐼𝜙𝜙̈+𝐺𝐺𝑋𝑋̇+𝐼𝐼𝜙𝜙̇𝜏𝜏⁄=−𝜅𝜅𝜆𝜆sin2𝜙𝜙+𝑇𝑇𝑚𝑚, (7) \nwhere 𝑀𝑀=2𝜌𝜌𝜌𝜌𝜆𝜆⁄, 𝐼𝐼=2𝜌𝜌𝜆𝜆𝜌𝜌, 𝐺𝐺=2𝛿𝛿𝑠𝑠𝜌𝜌, and 𝜏𝜏=𝜌𝜌/𝛼𝛼𝑠𝑠 are the mass, the moment of 6 \n inertia, the gyrotropic coefficient, and the relaxation time , respectively, and 𝜌𝜌 is the cross \nsectional area of the DW . Here, 𝐹𝐹𝑚𝑚=(2𝐴𝐴𝜆𝜆⁄)∫𝑑𝑑𝑑𝑑[(𝜕𝜕𝑥𝑥𝜌𝜌𝑚𝑚)𝒏𝒏𝟎𝟎×𝜕𝜕𝑖𝑖𝒏𝒏𝟎𝟎] and 𝑇𝑇𝑚𝑚=\n−2𝐴𝐴∫𝑑𝑑𝑑𝑑[(𝑱𝑱𝒎𝒎∙𝛁𝛁)𝒏𝒏𝟎𝟎] correspond to the magnon- induced force and torque , respectively. \nWe note that the sign of 𝑇𝑇𝑚𝑚 is different for left - and right -circular ly polarized SWs whereas \nthe sign of 𝐹𝐹𝑚𝑚 is independent of the circular polarization of SW. It is because the sign of 𝐽𝐽𝑚𝑚𝑥𝑥 \nis different for left - and right -circularly polarized SWs whereas the sign of 𝜕𝜕𝑥𝑥𝜌𝜌𝑚𝑚 is \nindependent of the circular polarization of SW. From Eqs. (6) and (7), we finally obtain the \nsteady -state velocity of DW below the Walker breakdown [ 53] as \n𝑣𝑣𝐷𝐷𝐷𝐷=𝑠𝑠\n2𝜌𝜌(𝛼𝛼2𝑠𝑠2+𝛿𝛿𝑠𝑠2)�𝛼𝛼𝜆𝜆𝐹𝐹𝑚𝑚+𝛿𝛿𝑠𝑠\n𝑠𝑠𝑇𝑇𝑚𝑚�, (8) \nwhich is the central result of this work. The first and second term s originate from non -\nadiabatic and adiabatic contributions , respectively. In Eq. (8), t he ratio 𝛿𝛿𝑠𝑠/𝑠𝑠 is an estimate of \nthe degree how the dynamics of the system is close to that of ferromagnets. The condition of \n𝛿𝛿𝑠𝑠2𝑠𝑠⁄→±1 represent s the ferromagnetic limit, whereas that of 𝛿𝛿𝑠𝑠2𝑠𝑠⁄→0 represent s the \nantiferromagnet ic limit. In the ferromagnetic limit (𝛿𝛿𝑠𝑠2𝑠𝑠⁄→±1), the second term in Eq. (8) \nbecomes dominant for the DW motion. On the other hand, in the antiferromagnetic limit \n(𝛿𝛿𝑠𝑠2𝑠𝑠⁄→0), the second term vanishe s and the only first term is responsible for DW motion . \nTo verify Eq. (8), we perform micromagnetic simulation s with the atomistic LLG \nequation. We start with the initial condition that the DW is located at the center of one -\ndimensional nanowire as shown in Fig. 1(b). SW is excited by an external AC field 𝐁𝐁𝐀𝐀𝐀𝐀 on \nthe left side of DW . The atomistic LLG equation including the external AC field is given by \n𝜕𝜕𝑺𝑺𝒊𝒊\n𝜕𝜕𝑡𝑡=−𝛾𝛾𝑖𝑖𝑺𝑺𝒊𝒊×�𝐁𝐁𝐞𝐞𝐞𝐞𝐞𝐞,𝒊𝒊+𝐁𝐁𝐀𝐀𝐀𝐀�+𝛼𝛼𝑖𝑖𝑺𝑺𝒊𝒊×𝜕𝜕𝑺𝑺𝒊𝒊\n𝜕𝜕𝑡𝑡, (9) \nwhere 𝑺𝑺𝒊𝒊 is the normalized spin moment vector, 𝛾𝛾𝑖𝑖=𝑔𝑔𝑖𝑖𝜇𝜇𝐵𝐵ℏ⁄ is the gyromagnetic ratio, \n𝜇𝜇𝐵𝐵 is the Bohr magneton, and 𝛼𝛼𝑖𝑖 is the damping constant at a lattice site 𝑖𝑖. The o dd (even) \nsite 𝑖𝑖 corresponds to the TM (RE) element. 𝐁𝐁𝐞𝐞𝐞𝐞𝐞𝐞,𝒊𝒊=−1\n𝜇𝜇𝑖𝑖𝜕𝜕ℋ\n𝜕𝜕𝑺𝑺𝒊𝒊 is the effective field at each \nsite, where 𝜇𝜇𝑖𝑖 is the magnetic mome nt per atom , one dimensional discrete Hamiltonian \nℋ=𝐴𝐴𝑠𝑠𝑖𝑖𝑚𝑚∑𝑺𝑺𝒊𝒊∙𝑺𝑺𝒊𝒊+𝟏𝟏 𝑖𝑖 −𝐾𝐾𝑠𝑠𝑖𝑖𝑚𝑚∑(𝑺𝑺𝒊𝒊∙𝒛𝒛�)2\n𝑖𝑖 +𝜅𝜅𝑠𝑠𝑖𝑖𝑚𝑚∑(𝑺𝑺𝒊𝒊∙𝒙𝒙�)2\n𝑖𝑖 , 𝐴𝐴𝑠𝑠𝑖𝑖𝑚𝑚 and 𝐾𝐾𝑠𝑠𝑖𝑖𝑚𝑚 (𝜅𝜅𝑠𝑠𝑖𝑖𝑚𝑚) are the 7 \n exchange constant and easy (hard) axis anisotropy for simulations , respectively . To excite the \nSW, an external AC field 𝐁𝐁𝐀𝐀𝐀𝐀=B0[cos𝜔𝜔𝑡𝑡𝒙𝒙�±sin𝜔𝜔𝑡𝑡𝒚𝒚�] is applied on two cells at \n252 nm away from the DW. We use the following simulation parameters: 𝐴𝐴𝑠𝑠𝑖𝑖𝑚𝑚=\n1.64 meV , 𝐾𝐾𝑠𝑠𝑖𝑖𝑚𝑚=6.47 𝜇𝜇eV, 𝜅𝜅𝑠𝑠𝑖𝑖𝑚𝑚=0.02𝐾𝐾𝑠𝑠𝑖𝑖𝑚𝑚, B0=100 mT, the lattice constant is \n0.42 nm, and the Land é 𝑔𝑔–factor 𝑔𝑔𝑅𝑅𝑅𝑅=2 for rare earth and 𝑔𝑔𝑇𝑇𝑀𝑀=2.2 for transition \nmetal [54]. We consider the damping constant is uniform for all site s, i.e., 𝛼𝛼RE=𝛼𝛼TM=5×\n10−4 for simplicity. We use m agnetic moment s 𝑀𝑀RE and 𝑀𝑀TM as listed in TABLE 1. \n \nIII. Results and Discussion \nFigure 3(a) -(c) show the simulation results of DW velocity as a function of the SW \nfrequency. Figure 3(a) represents the results for the case below the angular momentum \ncompensation point 𝑇𝑇𝐴𝐴. As the SW gap is different for left - and right -circular ly polarized \nSWs [Fig. 2(a)], the threshold SW frequenc y for the DW motion is also different for left - and \nright -circular ly polarized SWs. An interesting observation for the DW motion is that the \nmoving direction of DW depends on the circular polarization of SW . Left- (Right -) circular ly \npolarized SW moves the DW toward s (away from ) the SW source. This bi -directional DW \nmotion is understood by the fact that l eft- and right-circular ly polarized SWs carry the \nangular momentum with opposite signs. When the SW passes through the DW, the angular \nmomentum of SW is transferred to the DW so that the DW moving direction depends on the \ncircular polarization of SW . This is directly related to the fact that the sign of 𝑇𝑇𝑚𝑚 is different \nfor left - and right -circular ly polarized SWs , whereas the sign of 𝐹𝐹𝑚𝑚 is independent of the \ncircular polarization of SW. Given that 𝑇𝑇𝑚𝑚 and 𝐹𝐹𝑚𝑚 in Eq. (8) respectively correspond to \ncontributions from the adiabatic and non- adiabatic magnon ic torques , the bi -directional DW \nmotion depending on the circular polarization of SW evidences that the adiabatic magnon ic \ntorque is dominant over the non- adiabatic one. \nSolid and dashed lines in Fig. 3(a) are calculated from Eq. (8), with 𝐹𝐹𝑚𝑚 and 𝑇𝑇𝑚𝑚 \nobtained from numerical calculation s. We find that the numerically obtained bi -directional \nbehavior (symbols) is reasonably described by Eq. (8) in high- frequency ranges. In low-\nfrequency ranges , a discrepancy between Eq. (8) and numerical results appears possibly 8 \n because of nonlinear effects , which are not captured by our current analytical models . The \nresults for the case above the angular momentum compensation point 𝑇𝑇𝐴𝐴 [Fig. 3(c) ] can be \nunderstood in a similar way . Contrary to the case below 𝑇𝑇𝐴𝐴, overall spin moment s in the \nsystem are reversed so that left - (right -) circular ly polarized SW makes DW move away from \n(toward s) the source. \nTo further elucidate SW -induced ferrimagnetic DW motion below and above 𝑇𝑇𝐴𝐴, we \ninvestigate the spin current 𝐽𝐽𝑆𝑆, which is defined as 𝑱𝑱𝒔𝒔=−𝐴𝐴〈𝒏𝒏×𝜕𝜕𝑥𝑥𝒏𝒏〉. Figure 4 shows the \nschematic of SW transmission through a DW (top panel) and the z component of the spin \ncurrent 𝐽𝐽𝑆𝑆𝑧𝑧 along the propagation direction (i.e., the 𝑥𝑥 axis, bottom panel). For the left - \ncircular ly polarized SW (solid line) , the spin current in the left domain part decreases \ngradually due to the damping. After the SW passes through the DW, the spin current abruptly \nflips its sign due to overall reversal of spin moments. The spin -current change is transferred \nto the DW, resulting in the DW motion. For the right -circular ly polarized SW (dashed line), \noverall sign of the spin current is reversed. It is the reason that the direction of DW \npropagation is the opposite for left - and right -circularly polarized SWs. The sign of the spin-\ncurrent change is the same below and above 𝑇𝑇𝐴𝐴, but 𝛿𝛿 𝑠𝑠 changes its sign [see Eq. (8)] \nbecause the spin directions in the domain part changes accordingly , which results in the sign \ndifference of the DW velocity below and above 𝑇𝑇𝐴𝐴. \nFor the case at the angular momentum compensation point 𝑇𝑇𝐴𝐴 (i.e., 𝛿𝛿𝑠𝑠=0), both \nleft- and right-circular ly polarized SWs drive the DW to the same direction ( i.e., toward s the \nSW source) as shown in Fig. 3(b) . We note that this DW moving direction at 𝑇𝑇𝐴𝐴 is the \nopposite to the direction of the DW motion induced by circularly polarized spin waves in true \nantiferromagnets [ 40,41]. In true antiferromagnets where the shape anisotropy is absent, \ncircular ly polarized SWs make the DW precess, which result s in the SW reflection. The \nreflected SWs transfer linear momentum to DW, and push the DW away from the SW source. \nFor the case at 𝑇𝑇𝐴𝐴 of ferrimagnets, however, the net magnetic moment is finite so that the \nshape anisotropy does not vanish. As a result, the DW experiences the hard- axis anisotropy, \nwhich prevents the DW precession. Therefore, the ferr imagnetic DW still serves as a \nreflect ionless potential called the Pöschl -Teller potential [ 55] and its motion is governed by \nthe force from the magnonic torque [ 𝛼𝛼𝜆𝜆𝐹𝐹𝑚𝑚 term in Eq. (8)]. As the sign of 𝐹𝐹𝑚𝑚 is 9 \n independent of the SW circular polarization, both left - and right -circular ly polarized SWs pull \nthe DW along the same direction , i.e., towards the SW source. This force -induced motion of \nthe ferrimagnetic DW toward the spin -wave source at 𝑇𝑇𝐴𝐴 is similar to the motion of the \nantiferromagnetic DW toward the spin -wave source for lin early -polarized spin waves \nreported in Ref. [40] . \nDependence of the DW velocity on the SW circular polarization is summarized in \nFig. 3(d), which shows the DW velocity as a function of the net spin density 𝛿𝛿𝑠𝑠 at a fixed \nSW frequency (𝜔𝜔 = 0.7 THz) . The sign of DW velocity depends not only on the circular \npolarization, but also on the sign of the net spin density. We note that the DW velocity is not \nzero at 𝑇𝑇𝐴𝐴 because the adiabatic and non- adiabatic contributions are not compensated at 𝑇𝑇𝐴𝐴. \n \nIV. Summary \nWe have investigate d the SW circular -polarity dependence of ferrimagnetic DW \ndynamics theoretically and numerically. We find that the DW moves along the opposite \ndirection depending on the circular polarization of SW. This bi -directional DW motion is \ncaused by the fact that the signs of the spin current and the angular momentum transferred to \nDW are opposite for left - and right -circular ly polarized SWs . The o verall tendency of DW \nmoving direction is reversed when the sign of the net spin density 𝛿𝛿𝑠𝑠 is reversed. At 𝑇𝑇𝐴𝐴 \nwhere the angular momentum vanishe s, the dissipative non-adiabatic magnonic torque is the \nmain driving force so that DW moves along the same direction (towards the SW source) \nregardless of the SW circular polarization . \nOur finding of bi-directional ferrimagnetic DW driven by SWs can be generalized to \nother ferrimagnetic topological excitations such as magnetic skyrmions and vortices. This bi-\ndirectionality of ferrimagnetic DW motion depending either on the SW circular polarization \nor on the sign of the net spin density will be useful for magnonic spintronics [ 56] because \nsuch bi -directional motion, which makes the device functionality versatile, can be realized \nwithout moving the location of a SW source. \n 10 \n Acknowledgement \nThis work was supported by the National Research Foundation of Korea (NRF) \n(Grants No. 2015M3D1A1070465 and No. 2017R1A2B2006119), the KIST Institutional \nProgram (Project No. 2V05750). S.K.K. was supported by the startup fund at the University \nof Missouri. J.X. was supported by the National Natural Science Foundation of China (Grants No. 11722430). \n 11 \n Table 1. Used magnetic moments MTM and MRE for transition metal and rare earth elements , \nrespectively , in simulation. Index 5 coincides with the angular momentum compensation \npoint TA. \nIndex 1 2 3 4 5 6 7 8 9 \n𝑀𝑀𝑇𝑇𝑀𝑀(kA/m) 460 455 450 445 440 435 430 425 420 \n𝑀𝑀𝑅𝑅𝑅𝑅(kA/m) 440 430 420 410 400 390 380 370 360 \n \n 12 \n \nFigure 1. (a) Illustration of left- and right -circular ly spin wave s in an antiferromagnetically \ncoupled ferrimagnet. (b) Schematic graphic of one -dimensional ferrimagnet ic nanowire with \na domain wall (DW). Domain wall is positioned at the center of nanowire . Spin wave is \nexcited by an external AC field ( 𝐁𝐁𝐀𝐀𝐀𝐀) on the left side (252 nm apart from DW ). \n \n13 \n \nFigure 2. Spin- wave dispersion relations (a) below TA, (b) at TA, and (c) above TA. Symbols \nrepresent numerical simulation results and lines represent Eq. (4). Solid (Open ) triangular \nsymbols correspond to left - (right -) circularly polarized spin wave . \n \n14 \n \nFigure 3. Calculated domain wall velocity results (a) below TA, (b) at TA, and (c) above TA \nwith various spin- wave frequencies . Symbols are the simulation results and lines are Eq. (8) \n(arbitrary unit). (d) Domain wall velocity as a function of the net spin density δs at a fixed \nfrequency ω = 0.7 THz . Negative δs corresponds to the case below TA. \n \n15 \n \nFigure 4. Schematic of transm itted SW (top) and the z component of the spin current Js along \nthe wire length. A DW is positioned at the atomic site i = 2000, and SW source is at i = \n1700. Assumed parameters are those with the index 3 (i.e., below TA) listed in the TABLE 1 \nand the SW frequency is 0.6 THz. \n \n16 \n Reference \n1. R. Wangness, Phys. Rev. 91, 1085 (1953). \n2. M. Binder et al., Phys. Rev. B 74, 134404 (2006). \n3. C. Stanciu, A. Kimel, F. Hansteen, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing, \nPhys. Rev. B 73, 220402 (2006). \n4. J. M. D. Coey, “ Magnetism and Magnetic Materials ”, Cambridge University Press \n(2010). \n5. C. D. Stanciu, F. Hansteen, A. V . Kimel, A. Tsukamot o, A. Itoh, and T. Rasing, Phys. \nRev. Lett. 99, 047601 (2007). \n6. I. Radu et al., Nature 472 , 205 (2011). \n7. S. Mangin et al ., Nat. Mater. 13, 286 (2014). \n8. W. Cheng, X . Li, H . Wang, X . Cheng, and X . Miao , AIP Advances 7, 056018 (2017) . \n9. S. Alebrand et al., Appl. Phys. Lett. 101, 162408 (2012) . \n10. K. Vahaplar et al., Phys. Rev. B 85, 104402 (2012) . \n11. T. A. Ostler et al ., Nat. Commun. 3, 666 (2012) . \n12. T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich, Nat. Nanotechnol. 11 , 231 \n(2016). \n13. T. Jungwirth, J. Sinova, A. Manchon, X. Marti, J. Wunderlich, and C. Felser, Nat. \nPhys. 14, 200 (2018). \n14. O. Gomonay, V . Baltz, A. Brataas, and Y . Tserkovnyak, Nat. Phys. 14, 213 (2018) . \n15. R. A. Duine, K.- J. Lee, S. S. P. Parkin, and M. D. Stiles, Nat. Phys. 14, 217 (2018). \n16. N. Roschewsky et al ., Appl. Phys. Lett. 109, 112403 (2016). \n17. J. Finley and L. Liu, Phys. Rev. Applied 6, 054001 (2006). \n18. K. Ueda, M. Mann, C.- F. Pai , A.-J. Tan, and G. S. D. Beach, Appl. Phys. Lett. 109, \n232403 (2016). 17 \n 19. R. Mischra, J. Yu, X. Qiu, M. Motapothula, T. Venkatesan, and H. Yang, Phys. Rev. \nLett. 118, 167201 (2017). \n20. K.-J. Kim et al ., Nat. Mater. 16, 1187 (2017). \n21. S.-H. Oh, S. K. Kim, D. -K. Lee , G. Go, K. -J. Kim, T. Ono, Y . Tserkovnyak, and K. -J. \nLee, Phys. Rev. B 96, 10047(R) (2017). \n22. L. Caretta et al ., Nat. Nanotechnol. 13, 1154 (2018). \n23. S. A. Siddiqui, J. Han, J. T. Finley, C. A. Ross, and L. Liu, Phys. Rev. Lett. 121, 057701 (2018). \n24. S.-H. Oh a nd K.- J. Lee, J. Magn. 23, 196 (2018). \n25. M. Tanaka, H. Kanazawa, S. Sumitomo, S. Honda, K. Mibu, and H. Awano, Appl. \nPhys. Express 8, 073002 (2015). \n26. S. K. Kim, K. -J. Lee, and Y . Tserkovnyak, Phys. Rev. B 95, 140404(R) (2017). \n27. S. Woo et al., Nat. Commun. 9, 959 (2018). \n28. Y . Hirata et al., Nat. Nanotechnol. 14, 232 (2019). \n29. D.-H. Kim et al ., Phys. Rev. L ett. 122, 127203 (2019). \n30. J. Yu et al., Nat. Mater. 18 , 29 (2019). \n31. T. Okuno et al ., arXiv:1903.03251 (2019). \n32. D.-S. Han, S.- K. Kim, J. -Y . Lee, S. J. Hermsdoerfer, H. Schultheiss, B. Leven, and B. \nHillebrands, Appl. Phys. Lett. 94, 112502 (2009). \n33. S.-M. Seo, H.- W. Lee, H. Kohno, and K.- J. Lee, Appl. Phys. Lett. 98, 012514 (2011). \n34. P. Yan, X. Wang, and X. Wang, Phys. Rev. Lett. 107, 177207 (2011). \n35. X.-G. Wang, G. -H. Guo, Y .- Z. Nie, G. -F. Zhang, and Z.- X. Li, Phys. Rev. B 86, \n054445 (2012). 18 \n 36. P. Yan, A. K amra, Y . Cao, and G. E. W. Bauer, Phys. Rev. B 88, 144413 (2013). \n37. S. K. Kim and Y . Tserkovnyak, Phys. Rev. B 92, 020410 (2015). \n38. W. Wang et al., Phys. Rev. Lett. 114 , 087203 (2015). \n39. K.-W. Kim et al ., Phys. Rev. Lett. 122, 147202 (2019). \n40. E. G. Tveten, A. Qaiumzadeh, and A. Brataas, Phys. Rev. Lett. 112 , 147204 (2014). \n41. S. K. Kim, Y . Tserkovnyak, and O. Tchernyshyov, Phys. Rev. B 90, 104406 (2014). \n42. J. Lan, W . Yu, and J. Xiao, Nat. Commun. 8, 178 (2017). \n43. W. Yu, J. Lan, and J. Xiao, Phys. Rev. B 98, 144422 (2018). \n44. S. Pikin, Sov. Phys. JETP 27, 995 (1968). \n45. A. Andreev and V . Marchenki, Zh. Eksp. Teor, Fiz. 70, 1522 (1976). \n46. E. Solano- Carrillo, R. Franco, and J. Si lva-Valencia, Solid State Commun. 150, 2061 \n(2010). \n47. A. F. Andreev and V . I. Marchenko, Soviet Physics Uspekhi 23, 21 (1980). \n48. A. Chiolero and D. Loss, Phys. Rev. B 56, 738 ( 1997). \n49. B. A. Ivanov and A. L. Sukstanskii, Solid State Commun. 50, 523 ( 1984). \n50. O. Tretiakov, D. Clarke, G.- W. Chern, Y . B. Bazaliy, and O. Tchernyshyov, Phys. Rev. \nLett. 100, 127204 (2008). \n51. L. D. Landau, Course of Theoretical Physics 8, 15 (1960). \n52. R. Khoshlahni, A. Qaiumzadeh, A. Bergman, and A. Brataas, Phys. Rev. B 99, \n054423 (2018). \n53. N. L. Schryer and L. R. Walker, J. Appl. Phys. 45 , 5406 (1974). \n54. J. Jensen and A. R. Mackintosh, Rare Earth Magnetism (Clarendon, Oxford, UK, 19 \n 1991). \n55. G. Pöschl and E. Teller, Zeitschrift für Physik 83, 143 (1933). \n56. A. V . Chumak, V . I. Vasyuchka, A. A. Serga, and B. Hillebrands, Nat. Phys. 11, 453 \n(2015). \n " }, { "title": "1005.5297v2.Haldane_Phases_and_Ferrimagnetic_Phases_with_Spontaneous_Translational_Symmetry_Breakdown_in_Distorted_Mixed_Diamond_Chains_with_Spins_1_and_1_2.pdf", "content": "arXiv:1005.5297v2 [cond-mat.str-el] 10 Nov 2010Typeset with jpsj3.cls Full Paper\nHaldane Phases and Ferrimagnetic Phases with Spontaneous T ranslational\nSymmetry Breakdown in Distorted Mixed Diamond Chains with S pins1and1/2\nKazuoHida∗, Ken’ichi Takano1, and Hidenori Suzuki1†\nDivision of Material Science, Graduate School of Science an d Engineering,\nSaitama University, Saitama, Saitama, 338-8570\n1Toyota Technological Institute, Tenpaku-ku, Nagoya 468-8 51\n(Received February 28, 2018)\nThe ground states of two types of distorted mixed diamond cha ins with spins 1 and 1 /2\nare investigated using exact diagonalization, DMRG, and ma pping onto low-energy effective\nmodels. In the undistorted case, the ground state consists o f an array of independent spin-\n1 clusters separated by singlet dimers. The lattice distort ion induces an effective interaction\nbetween cluster spins. When this effective interaction is an tiferromagnetic, several Haldane\nphases appear with or without spontaneous translational sy mmetry breakdown (STSB). The\ntransition between the Haldane phase without STSB and that w ith (n+1)-fold STSB ( n= 1,\n2, and 3) belongs to the same universality class as the ( n+1)-clock model. In contrast, when\nthe effective interaction is ferromagnetic, the quantized a nd partial ferrimagnetic phases appear\nwith or without STSB. An effective low-energy theory for the p artial ferrimagnetic phase is\npresented.\nKEYWORDS: mixed diamondchain, distortion, frustration, H aldane phase, spontaneoustranslational sym-\nmetry breakdown, partial ferrimagnetism\n1. Introduction\nQuantum magnetism in frustrated spin systems is a\nrapidly developing field of condensed matter physics.1,2)\nAt first glance, one would expect that geometrical frus-\ntration enhances quantum fluctuation and drives an or-\ndered state into a disordered state. However, recent\nprogress in this field of physics has shown that this sim-\nple intuition is not always valid and that geometrical\nfrustration induces a variety of exotic quantum phenom-\nena,whicharenoteasilypredicted.Underanappropriate\ncondition, it even stabilizes an unexpected magnetic long\nrangeordersuch asthe frustration-inducedferrimagnetic\nand spin nematic orders.\nTo understand magnetism under the interplay of geo-\nmetrical frustration and quantum fluctuation, it is desir-\nable to begin with typical spin models with exact solu-\ntions. Among them, there exist a class of models whose\nground states are exactly written down as spin cluster\nsolid (SCS) states because of frustration. A SCS state is\natensorproduct stateofexactlocaleigenstatesofcluster\nspins. Well-known examples are the Majumdar-Ghosh\nmodel3)whose ground state is a prototype of sponta-\nneously dimerized phases in one-dimensional frustrated\nmagnets4)and the Shastry-Sutherland model5)which\ncorresponds to the material SrCu 2(BO3)2.6,7)In these\nmodels, the spin clusters are singlet dimers.\nThe diamond chain is another frustrated spin chain\nwith exact SCS ground states. The lattice structure is\nshown in Fig. 1. In a unit cell, there are two kinds of\nnonequivalent lattice sites occupied by spins with mag-\nnitudesSandτ; we denote the set of magnitudes by ( S,\nτ). One of the authors and coworkers8,9)introduced this\n∗E-mail address: hida@phy.saitama-u.ac.jp\n†Present address: Department of Physics, College of Humanit ies\nand Sciences, Nihon University, Setagaya-ku, Tokyo 156-85 50Sτ\nτ\nFig. 1. Structure ofthe diamond chain. Spin magnitudes ina u nit\ncell are indicated by Sandτ; we denote the set of magnitudes\nby (S,τ). The PDC is the case of S=τ, while the MDC is the\ncase ofS= 2τwith an integer or half-odd integer τ.\nlattice structure and generally investigated the case of\n(S,S), i.e., the pure diamond chain (PDC). Any PDC\nis shown to have at least one exact SCS ground-state\nphase where each spin cluster has spin 0. Particularly,\nin the case of (1/2, 1/2), they determined the full phase\ndiagram of the ground state by combining rigorous ar-\nguments with numerical calculations. After that, Nigge-\nmann et al.10,11)arguedabout a seriesofdiamond chains\nwith (S, 1/2). As for the special case of (1/2, 1/2), they\nreproduced the results of ref. 9.\nThe mixed diamond chain (MDC) is defined as a dia-\nmond chain with ( S,S/2) for the integer S.12)The spe-\ncial case of (1, 1/2) was first investigated by Niggemann\net al.10,11)They considered it asone of the seriesofmod-\nels with ( S,1/2).Recently, extensiveinvestigationonthe\nMDC has been carried out by the present authors.12–14)\nThe MDC is of special interest among diamond chains,\nbecause only the MDC has the Haldane phase in the ab-\nsence offrustration,so that wecan observethe transition\nfrom the Haldane phase to a SCS phase induced by frus-\ntration. In contrast, diamond chains of other types have\nferrimagnetic ground states for weak frustration.\nThe features common to all types of diamond chains\nare their infinite number of local conservation laws and\n12 J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki\nmore than two different types ofexact SCS groundstates\nthat are realized depending on the strength of frustra-\ntion. For example, S= 1/2 PDC has a nonmagnetic\nphaseaccompaniedbyspontaneoustranslationalsymme-\ntry breakdown (STSB) and a paramagnetic phase with-\nout STSB. This model also has a ferrimagnetic ground\nstate in the less frustrated region.9)On the other hand,\nthe MDC with spins 1 and 1 /2 has 3 different param-\nagnetic phases accompanied by STSB and one paramag-\nnetic phase without STSB. This model also has a non-\nmagnetic Haldane ground state in a less frustrated re-\ngion.10,12)The SCS structures of the ground states are\nalso reflected in characteristic thermal properties, as re-\nported in ref. 13.\nModifications of the PDC and MDC have been exam-\nined by many authors. Among them, the spin 1/2 PDC\nwith distortion has been thoroughly investigated by nu-\nmerical methods.15–17)It is found that azurite, a nat-\nural mineral, consists of distorted PDCs with spin 1/2\nand that the magnetic properties of this material have\nbeen experimentally studied in detail.18,19)Other mate-\nrials have also been reported.20,21)The diamond chain\nis one of the simplest models compatible with the 4-spin\ncyclic interaction. The effects of this type of interaction\non PDC have recently been investigated by Ivanov et\nal.22)The present authors also investigated the MDC\nwith bond-alternating distortion and found an infinite\nseries of ground states with STSB.14)In addition, as re-\nviewed in ref. 14, the MDC is related to other impor-\ntant models of frustrated magnetism such as the dimer-\nplaquette model,23–28)frustrated Heisenberg ladders,29)\nhybrid diamond chains consisting of Heisenberg bonds\nand Ising bonds,30,31)and an Ising model on a hierarchi-\ncaldiamondlattice.32)Amongthem,thedimer-plaquette\nchain with ferromagnetic interplaquette interaction re-\nduces to the MDC in the limit of strong interplaquette\ninteraction.28)\nThus far, in spite of the theoretical relevance of the\nMDC, no materials described by the MDC have been\nfound. Nevertheless, synthesizing MDC materials is not\nan unrealistic expectation in view of the success of the\nsynthesis of many low dimensional bimetallic magnetic\ncompounds33)and organic magnetic compounds.34)In\ngeneral, it is natural to expect that the lattice is pos-\nsibly distorted in real MDC compounds as in azurite.\nFrom this viewpoint, it is important to present theoret-\nical predictions on the ground state of distorted MDCs\nto widen the range of candidate materials of MDC and\nto raise the possibility of their synthesis.\nWe begin by classifying the distortion patterns by\nthe normal modes of each diamond unit. Excluding two\ntranslations and one rigid body rotation, we have 5 nor-\nmal modes as depicted in Fig. 2 within the diamond\nplane. A distorted MDC may be realized as a result of\nthe collective softening of these normal modes. In partic-\nular, the distortion patterns in (a) and (b) break the lo-\ncal conservation laws that hold in the undistorted MDC.\nHence, these distortions induce effective interactions be-\ntween the cluster spins in the whole lattice, and may\nform novel exotic phases. We investigate these interest-\ning cases in the present paper. In what follows, we name(a) (b) (c)\n(d) (e)\nFig. 2. Displacement modes of a diamond unit.\nSl1+δA\nλ\n1+δA1−δA\n1−δAτl(1)\nSl+1\nτl(2)(a)\nSl1+δB\nλ\n1−δB1+δB\n1−δBτl(1)\nSl+1\nτl(2)(b)\nFig. 3. Structures of MDC with S= 1 and τ(1)=τ(2)= 1/2\nwith (a) type A and (b) type B distortions.\nthe distortion patterns in (a) and (b) as type A and type\nB, respectively. The MDCs with type A and type B dis-\ntortions are depicted in Figs. 3(a) and 3(b), respectively.\nThe distortion patterns in Figs. 2(d) and 2(e) do not\nchange the geometry of the original undistorted MDC.\nThe distortion pattern in Fig. 2(c) is of another interest,\nsince it induces the bond alternation in the undistorted\nMDC without breaking the local conservation laws. This\ncase has been investigated separately and published in a\nprevious paper.14)\nThis paper is organized as follows. In §2, the Hamil-\ntonians for the MDCs with the type A and type B dis-\ntortions are presented, and the structure of the ground\nstatesoftheMDCwithoutdistortionissummarized.The\nground-statephasesfor the MDC with the type Adistor-\ntion are discussed in §3, and those for the MDC with the\ntype B distortion are discussed in §4. The last section is\ndevoted to summary and discussion.\n2. Hamiltonian\nThe MDCs with the type A and type B distortions are\ndescribed, respectively, by the following Hamiltonians:\nHA=N/summationdisplay\nl=1/bracketleftBig\n(1+δA)Slτ(1)\nl+(1−δA)τ(1)\nlSl+1\n+(1−δA)Slτ(2)\nl+(1+δA)τ(2)\nlSl+1+λτ(1)\nlτ(2)\nl/bracketrightBig\n,\n(1)J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki 3\nHB=N/summationdisplay\nl=1/bracketleftBig\n(1+δB)Slτ(1)\nl+(1+δB)τ(1)\nlSl+1\n+(1−δB)Slτ(2)\nl+(1−δB)τ(2)\nlSl+1+λτ(1)\nlτ(2)\nl/bracketrightBig\n,\n(2)\nwhereSlis the spin-1 operator, and τ(1)\nlandτ(2)\nlare\nthe spin-1/2 operators in the lth unit cell. The param-\neterδA(δB) represents the strength of type A (type B)\ndistortion, and is taken to be nonnegative without spoil-\ning generality. The number of unit cells is denoted by\nN, and then the total number of sites is 3 N. We will\nconsider these systems in the large Nlimit.\nForδA= 0 and δB= 0, both eqs. (1) and (2) reduce\nto the undistorted MDC Hamiltonian,\nH0=N/summationdisplay\nl=1/bracketleftbigg\nSlTl+TlSl+1+λ\n2/parenleftbigg\nT2\nl−3\n2/parenrightbigg/bracketrightbigg\n(3)\nwith the composite spin operators Tl≡τ(1)\nl+τ(2)\nl.\nBefore going into the analysis of the distorted MDC,\nwe briefly summarize the ground-state properties of the\nHamiltonian (3) reported in ref. 12 for convenience.\n(i)T2\nlcommutes with the Hamiltonian H0for anyl.\nTherefore, the composite spin magnitude Tldefined\nbyT2\nl=Tl(Tl+1) is a good quantum number that\ntakes the values 0 or 1. Hence, each energy eigen-\nstate has a definite set of {Tl}, i.e. a sequence of\n0’s and 1’s with length N. A pair of τ(1)\nlandτ(2)\nl\nwithTl= 0 is called a dimer. A cluster including\nnsuccessive Tl= 1 pairs bounded by two Tl= 0\npairs is called a cluster- n. The cluster- nis equiva-\nlent to an antiferromagneticspin-1 Heisenberg chain\noflength2 n+1withopen boundarycondition.Since\na cluster- nis decomposed into a sublattice consist-\ning ofn+1 sites with Sl’s and that consisting of n\nsites with Tl’s, the ground states of a cluster- nare\nspin triplet states with total spin unity on the basis\nof the Lieb-Mattis theorem.35,36)This implies that\neach cluster- ncarries a spin-1 in its ground state.\n(ii) There appear 5 distinct ground-state phases called\ndimer-cluster- n(DCn) phases with n= 0,1,2,3,\nand∞. The DC nstate is an alternating array\nof dimers and cluster- n’s. The phase boundary\nλc(n,n′) between DC nand DCn′phases are\nλc(0,1) = 3,\nλc(1,2)≃2.660425045542 ,\nλc(2,3)≃2.58274585704 ,\nλc(3,∞)≃2.5773403291 , (4)\nwhereλc(0,1)is obtainedanalyticallyandotherval-\nues are calculated numerically.\n(iii) In the DC ∞ground state realized for λ < λc(3,∞),\nTl= 1 for all l. This state is not accompanied by\nSTSB and is equivalent to the Haldane state of an\nantiferromagnetic spin-1 Heisenberg chain with infi-\nnite length.\n(iv) Each of the DC nstates with 0 ≤n≤3 realized forλ > λ c(3,∞) is a uniform array of cluster- n’s with\na common value of nand dimers in between. In the\nDCnphase with 1 ≤n≤3, (n+1)-fold STSB takes\nplace. In the DC0 phase, no translational symmetry\nis broken.\nIn what follows, we numerically examine various aspects\nof the type A and type B distortion effects on the MDC.\nBecause the DC3 phase is only realized within a very\nnarrow interval of λ, it is difficult to analyze the effect\nof distortion numerically in this phase. Hence, we do not\nconsider the DC3 phase in the following numerical anal-\nysis.\n3. Ground-State Properties of the MDC with\nType A Distortion\n3.1 Weak distortion regime\nWe now inspect the nature of the effective interaction\nbetween two cluster- n’s separated by a dimer consisting\nofτ(1)\nlandτ(2)\nlin the presence of the weak type A dis-\ntortion. For δA>0,Sl(Sl+1) tends to be antiparallel to\nτ(1)\nl(τ(2)\nl) rather than to τ(2)\nl(τ(1)\nl), as is known from\nFig. 3(a). The spins τ(1)\nlandτ(2)\nlare antiparallel to each\nother because they form a singlet dimer. Therefore, Sl\nandSl+1tend to be antiparallel to each other. In each\ncluster-n, the number of spins Sl’s is larger than the\nnumber of composite spins Tl’s by one. Hence, from the\nLieb-Mattistheorem,35)thetotalspinofthegroundstate\nof the cluster- npoints to the same direction as the Sl’s\nbelonging to that cluster- n. Therefore, the total spins of\nthe cluster- n’s on both sides of the dimer also tend to be\nantiparallel to each other. Thus, the effective coupling\nbetween the spins of neighboring cluster- n’s is antifer-\nromagnetic. This physical argument will be numerically\nensured below.\nIn general, the interaction between two spins with a\nmagnitude of 1 is the sum of bilinear and biquadratic\nterms. Therefore, the effective Hamiltonian for cluster-\nn’s in the phase that continues to the DC nphase in the\nlimit ofδA→0 is written as\nHeff=Nc/summationdisplay\ni=1Heff(i,i+1), (5)\nHeff(i,i+1) =JeffˆSiˆSi+1+Keff/parenleftBig\nˆSiˆSi+1/parenrightBig2\n,(6)\nwhereˆSiis the total spin of the i-th cluster- nwith a\nmagnitude of 1, Ncis the total number of cluster- n’s,\nandJeffandKeffare effective coupling constants. From\nsymmetry consideration, the signs of δAdoes not affect\nthe sign of the effective coupling constants. Hence, these\ncoupling constants are of the order of δ2\nAfor small δA.\nWe numerically calculated the ground-state energy of a\npair of cluster- n’s with total spin Stot, and compared\nit with the corresponding eigenvalues of Heff(i,i+ 1).\nThen we confirmed that Jeff/δ2\nAandKeff/δ2\nAare almost\nindependent of δAtypically for δA<0.002. The constant\nvaluesof Jeff/δ2\nAandKeff/δ2\nAareshowninFig.4forthree\nphases (n=0, 1, and 2), which will be explained below.\nBecause the effective coupling constants satisfy 0 <\nKeff/Jeff<1, the ground state is the Haldane state for4 J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki\n3 4 50510\n0.51\nλ λc(0,1)Keff/Jeff\nKeff/δA2Jeff/δA2Keff/JeffHDC0 Jeff/δA2\nKeff/δA2\n2.7 2.8 2.9 302040\n00.51\nλλc(0,1) λc(1,2)Keff/Jeff\nKeff/δA2Jeff/δA2Keff/Jeff HDC1Jeff/δA2\nKeff/δA2\n2.6 2.62 2.64 2.66050100\n00.51\nλc(1,2) λc(2,3)Keff/Jeff\nKeff/δA2Jeff/δA2Keff/Jeff HDC2 Jeff/δA2\nKeff/δA2\nλ\nFig. 4. Effective bilinear interaction ( Jeff) and biquadratic inter-\naction (Keff) between spin clusters for small δAin HDC0, HDC1,\nand HDC2 phases from top to bottom. The ratio Keff/Jeffis also\nshown.\nsmallδA.37)In the Haldane state, each spin-1 degree of\nfreedom is carried by a cluster- nrather than by a sin-\ngle spin. We call the state the Haldane DC n(HDCn)\nstate. In the HDC nstate with n≥1, the (n+ 1)-fold\ntranslational symmetry is spontaneously broken unlike\nthe conventional Haldane state without STSB. Both the\nHDC0 state for λ > λ c(0,1) and the HDC ∞state for\nλ < λ c(3,∞) are the Haldane states without STSB. In\nparticular, the HDC ∞state continues from the Haldane\nstate (DC ∞state) of the undistorted MDC mentioned\nin§2.12)Uniform Haldane\nλ>λc(1,0)\nHDC1\nHDC2\nHDC3Uniform Haldane\nλ>λc(1,0)\nUniform Haldane\nλ<λc(3,∞) or δA » 1(a)\n(b)\n(c)\n(d)\n(e)\nFig. 5. Valence bond structures of the ground states of all ph ases\nfor the MDC with type A distortion. A small filled circle repre -\nsents a spin with a magnitude of 1/2. An original spin with a\nmagnitude of 1 is represented by two decomposed 1/2 spins in\nan open circle indicating the symmetrization. A valence bon d is\nrepresented by a dashed oval.\n3.2 Connection to the strong distortion regime\nIn the strong distortion regime of δA→1 and small\nλ, the three spins τ(2)\nl−1,Sl, andτ(1)\nlform a singlet clus-\nter. Hence, the ground state is a state with spin gap and\nwithout STSB. This nature is common to the HDC0 and\nHDC∞phases in §3.1. Furthermore, the HDC ∞state\nis transformed into the HDC0 state only by rearranging\ntwo valence bonds within each diamond unit, as shown\nin Figs. 5(a) and 5(e). Therefore, the strong distortion,\nHDC0, and HDC ∞regimes are considered to be differ-\nent parts of a single phase. The continuity of the three\nregimes will be confirmed by the numerical analysis dis-\ncussed in §3.3. In what follows, we call this phase the\nuniform Haldane (UH) phase as a whole.\n3.3 Numerical phase diagram\nUnder the periodic boundary condition, even in the\nparameter region where STSB takes place, the ground\nstate of a finite chain is a superposition of the symmetry-\nbroken states, and the translational symmetry is re-\ncovered. Under the open boundary condition, however,\none of the symmetry broken states is selected by the\nboundary effect. Therefore, we employ the DMRG cal-\nculation with the open boundary condition to determine\nthe phase diagram for finite δA. The DMRG calculation\nis carried out using the finite-size algorithm up to 288\nsites keeping 200 states in subsystems. We calculate the\nground-state expectation values/angbracketleftbig\nT2\nl/angbracketrightbig\nand define the ef-\nfective spin magnitude Tlon thel-th diagonal bond by\nTl(Tl+1) =/angbracketleftbig\nT2\nl/angbracketrightbig\n. A typical ldependence of Tlis shown\ninFig.6ineachphase.Withtheincreasein δA,thetrans-\nlational symmetry is recoveredas expected. For finite δA,\nthe ground-state phase is identified from the periodicity\nin the oscillation of Tl. In the HDC nphase, the values ofJ. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki 5\n0 5001\nlTl\nδA=0.04\nδA=0.02λ=2.803N=288(a)\n0 5001\nlTl\nδA=0.009\nδA=0.007λ=2.623N=282(b)\nFig. 6. Profiles of Tlfor (a)λ= 2.8 and (b) λ= 2.62.\n0 0.2 0.4 0.600.20.40.6∆T\nN−1/8δA=0.0304\nδA=0.0306\nδA=0.0308(a)\n0 0.2 0.4 0.600.10.2∆T\nN−2/15δA=0.00824δA=0.00822\nδA=0.00826(b)\nFig. 7. System size dependences of ∆ Tat (a)λ= 2.85 and (b)\nλ= 2.62. The data are plotted against N−β/νwhereβandν\nare the critical exponents of the order parameter and correl ation\nlength, respectively, for the 2-dimensional (a) Ising and ( b) 3-\nclock model.\nTlfollow the sequence\n...TSTL···TL/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright\nnTS,TL···TL/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright\nn..,(TL> TS).(7)\nThus, we define the order parameter of the HDC nphase\nby ∆T=TL−TS. In DMRG, ∆ Tis measured at the\nsites closest to the center of the chain.\nThe valence bond structures for the HDC nphases as2.6 2.8 300.020.04\nδA\nλλc(2,3)λc(3,∞)Z2STSB\n Z3STSBNo STSB\nλc(1,2) λc(0,1)HDC1\nHDC2Uniform Haldane\nFig. 8. Phase diagram of the MDC with type A distortion. The\ntriangles indicate the position of the phase boundary for δA= 0.\nwell as the UH phase are shown in Fig. 5. We see the\ntranslational invariance of period n+ 1 in the HDC n\nground state in contrast to the period-1 invariance in\nthe UH ground state. Hence, the Zn+1STSB takes place\nat the HDC n-UH phase boundary. We expect that this\ntransition belongs to the 2-dimensional ( n+ 1)-clock\nmodel universality class. The system size dependence\nof ∆Tforλ= 2.85 is shown in Fig. 7(a) around the\nHDC1-UH phase boundary. Here, the data are plotted\nagainstN−β/νwith the order parameter critical expo-\nnentβ= 1/8 and the correlation length critical expo-\nnentν= 1 for the two-dimensional Ising universality\nclass.This showsthatthecriticalvalueof δAliesbetween\n0.0304 and 0.0308. A similar plot is shown in Fig. 7(b)\nforλ= 2.62 around the HDC2-UH phase boundary, as-\nsuming the critical exponents of two-dimensional 3-clock\nmodel(equivalently3-statePottsmodel38))withβ= 1/9\nandν= 5/6. This shows that the critical value of δA\nlies between 0.00822 and 0.00826. The critical points at\nother values of λare determined similarly. The results\nare shown in the phase diagram of Fig. 8. The error bars\nare within the size of the symbols.\nTo confirm the consistency of the universality class,\nthe finite size scaling plot for the order parameter ∆ Tis\ncarried out. According to the scaling hypothesis, the δA\ndependence of the order parameter ∆ Tof the finite size\nsystems near the critical point should obey the finite size\nscaling law39)\n∆TNβ/ν=f(N(δA−δc\nA)ν), (8)\nintermsofthescaledvariables∆ TNβ/νandN(δA−δc\nA)ν\nand the scaling function f(x). In Figs. 9(a) and 9(b),\n∆TNβ/νis plotted against N(δA−δc\nA)νaround the\nHDC1-UH and HDC2-UH phase boundaries assuming\nthe Ising and 3-clock universality classes, respectively.\nThe critical points δc\nA= 0.0307 (Fig. 9(a)) and 0.008248\n(Fig. 9(b)) are chosen so that all data fall on a single\nuniversal scaling curve as well as possible. These plots\nare consistent with the expected universality class.\nThe critical behavior at the HDC1-UH transition in6 J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki\n0 0.101∆TN1/8\n(δA−δAc)N3N=144\n216\n288(a)\n−0.2 0 0.201∆TN2/15\n(δA−δAc)5/6N3N=138\n210\n282(b)\nFig. 9. Finite-size scaling plot of ∆ Taround the critical points.\n(a) Plot around the HDC1-UH phase boundary at λ= 2.85.\nThe Ising critical exponents ν= 1 and β= 1/8 are assumed.\nThe critical point is set at δc\nA= 0.0307. (b) Plot around the\nHDC2-UHphase boundary at λ= 2.62. The3-state Potts critical\nexponents ν= 5/6 andβ= 1/938)are assumed. The critical\npoint is set at δc\nA= 0.008248.\n(a)\n(b)\nFig. 10. Valence bond structures of the ground states of spin -1\nbilinear-biquadraticchaininthe (a)Haldanephase and (b) dimer\nphase. The spins with a magnitude of unity represented by ope n\ncircles are decomposed into two spin-1/2 degrees of freedom rep-\nresented by smallfilled circles.The valence bonds are repre sented\nby dashed ovals. The spins belonging to disconnected cluste rs in\nthe dimer phase are connected by the valence bonds in the Hal-\ndane phase.\nour model should be compared with that of the S=\n1 bilinear-biquadratic chain at the Takhtajan-Babujian\npoint.40,41)Both transitions are accompanied by Z2-\nSTSB which contributes to the conformal charge by 1/2.\nFortheHDC1-UHtransitioninourmodel,therearrange-\nment of valence bonds take place only within each dia-\nmond unit, as shown in Fig. 5(a) and 5(b). In contrast,\nin theS= 1 bilinear-biquadratic chain, the spins be-\nlonging to disconnected clusters in the dimer phase are0 1 2 300.51\nλδB 3N=18\nUHFDC0\nm=1\nFDC1\nm=1/2\nFDC2\nm=1/3\nFig. 11. Phase diagram of the MDC with type B distortion with\n3N= 18. The triangles indicate the position of the phase bound-\nariesλc(n,n+1) for δB= 0.12)\nconnected by the valence bonds in the Haldane phase, as\nshown in Fig. 10. Apart from Z2-STSB, this is similar\nto the Gaussian criticality of the Haldane-dimer transi-\ntion in the spin-1 alternatingbond Heisenbergchain that\ncontribute to the conformal charge by 1.42–44)Therefore,\ntheS= 1 bilinear-biquadratic chain at the Takhtajan-\nBabujian point is described by the conformal field theory\nwithc= 1/2+1 = 3 /2, while the HDC1-UH transition\nin our model is described by the c= 1/2 Ising conformal\nfield theory.\n4. Ground-State Properties of the MDC with\nType B Distortion\nIn the case of type B distortion, the effective inter-\naction between the spins of two cluster- n’s separated by\nthedimerconsistingof τ(1)\nlandτ(2)\nlisferromagnetic,be-\ncause both SlandSl+1tend to be antiparallel to τ(1)\nl.\nTherefore,we expect the ferrimagneticgroundstate with\nspontaneous magnetization quantized as m= 1/(n+1)\nper unit cell for small δBin the range λc(n,n+ 1)<\nλ < λc(n−1,n). We call this phase a ferrimagnetic DC n\nphase (FDC nphase). In contrast, the ground state for\nλ < λ c(3,∞) will remain in the Haldane phase, since a\nnonmagnetic gapped phase is generally robust against a\nweakdistortion.Forfinite δB,we determined the ground-\nstate phase diagramby the numerical diagonalizationfor\nthe system size 3 N= 18, as shown in Fig. 11. Among\nsystem sizes tractable by numerical diagonalization, only\nthis size of 3 N= 18 is compatible with all the ground-\nstate structures with n= 0,1, and 2. As expected, the\nFDCnphases with m= 1/(n+ 1) are found for these\nvalues of n.\nBy inspecting numerical data for the 3 N= 18 system,\nwe also find other narrow ferrimagnetic phases between\nthe FDC nand FDC( n+ 1) phases with n= 0, 1, and\n2, although they are too narrow to be shown in Fig. 11.\nIn order to investigate these phases in detail, we employJ. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki 7\n2.4 2.6 2.800.51\nλm\n3N=18(ED)\n3N=72(DMRG)δB=0.2(a)\n1.7 1.800.51\nλm\nλm\n3N=18(ED)\n3N=72(DMRG)δB=0.6(b)\nFig. 12. Spontaneous magnetization for (a) δB= 0.2 and (b)\nδB= 0.6. The triangles on the vertical axes indicate the values\nof the spontaneous magnetization m= 1/(n+ 1) in the FDC n\nphases.\nthe DMRG calculation for 3 N= 72 keeping 120 states in\neach subsystem. Typical examples of the λdependence\nof spontaneous magnetization are shown in Fig. 12(a)\nforδB= 0.2 and Fig. 12(b) for δB= 0.6. Between the\nFDCnand FDC( n+1) phases with n= 0,1,2, we find\nthe partial ferrimagnetic phase in which the spontaneous\nmagnetization varies continuously with λ. The ferrimag-\nnetic phase of this kind has been found in various frus-\ntrated one-dimensional quantum spin systems.45–51)In\ncontrast, between the nonmagnetic phase and the FDC3\nphase, we find no partial ferrimagnetic phase for small\nδB.\nThis can be understood as follows: At λ=λc(n,n+1),\nthe cluster- nand cluster-( n+1) can coexist. As stated\nabove, it is physically evident that the effective magnetic\ninteraction between the clusters is ferromagnetic. There-\nfore, wecan restrict the statesof eachcluster to the max-\nimally polarized ground state with ˆSz\ni= 1. Hence, the\ngroundstateofthe wholechainisdescribedbyspecifying\nthe arrangement of cluster- n’s and cluster-( n+1)’s. We\nmap the two possible values of the length of i-th cluster,\nni=nandni=n+ 1, to two possible values of the\nspin-1/2 pseudospin, σz\ni= 1/2 andσz\ni=−1/2, respec-\ntively. Then, the total magnetization Mis equal to the\nnumber of clusters Nc. The total number of unit cells,N, is related to the pseudospins σz\nias\nN=Nc/summationdisplay\ni=1/parenleftbigg\nn+1+1\n2−σz\ni/parenrightbigg\n=Nc/parenleftbigg\nn+3\n2/parenrightbigg\n−Nc/summationdisplay\ni=1σz\ni.\n(9)\nTherefore, the ground-state magnetization per unit cell\nmis given by\nm=Nc\n/an}bracketle{tN/an}bracketri}ht=1\nn+3\n2−σ(10)\nwithσ≡/summationtextNc\ni=1/an}bracketle{tσz\ni/an}bracketri}ht/Nc.Thebracket /an}bracketle{t···/an}bracketri}htrepresentsthe\nground-state expectation value. In the presence of δB,\nthe length of neighboring clusters can exchange through\na second order process in δB. This corresponds to the\nspin exchange in terms of pseudospins. In this case, the\ninteraction between the pseudospins is approximated by\nthe spin-1 /2 XXZ Hamiltonian\nHXXZ=Nc/summationdisplay\ni=1HXXZ(i,i+1), (11)\nHXXZ(i,i+1) =Jeff\nzσz\niσz\ni+1+Jeff\n⊥(σx\niσx\ni+1+σy\niσy\ni+1)\n(12)\nuptothesecondorderin δB.Here,furtherneighborinter-\nactions are neglected. We estimate the effective exchange\nconstants by comparing the energy spectrum of the pair\nHamiltonian HXXZ(i,i+1) with that of the correspond-\ning pair of clusters as follows:\n(i)λ=λc(0,1)\nJeff\nz≃ −0.039δB2,\nJeff\n⊥≃0.087δB2. (13)\n(ii)λ=λc(1,2)\nJeff\nz≃ −0.0082δB2,\nJeff\n⊥≃0.069δB2. (14)\n(iii)λ=λc(2,3)\nJeff\nz≃ −0.0029δB2,\nJeff\n⊥≃0.018δB2. (15)\nThe details of the calculation are explained in Appendix.\nIn all cases (i) ∼(iii), we find that the effective cou-\npling constants satisfy the inequality −|Jeff\n⊥|< Jeff\nz≤\n|Jeff\n⊥|. As is well known, the ground state of the spin-\n1/2 XXZ chain in this parameter regime is nonmagnetic\nand gapless in the absence of a magnetic field. Roughly\nspeaking, ∆ λ≡λ−λc(n,n+1) corresponds to the effec-\ntive magnetic field heffconjugate to the total pseudospin/summationtext\niσz\ni, because the increase in λfavors cluster- nover\ncluster-(n+1); however, this correspondence should not\nbe taken literally. A more precise argument is also given\nin Appendix. When ∆ λtakes a large negative value,\nthe pseudospins are fully polarized downward to give\n/an}bracketle{tσz\nl/an}bracketri}ht=−1/2. This state corresponds to the FDC( n+1)\nstate with m= 1/(n+2). When heffreaches the critical\nvaluehc1≡ −(|Jeff\n⊥|+Jeff\nz), the magnetization starts to\nincrease continuously until all spins are fully polarized8 J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki\nupward at the critical effective field hc2≡ |Jeff\n⊥|+Jeff\nz.\nThis corresponds to the FDC nstate with m= 1/(n+1).\nOn the other hand, the magnetization jumps from 0 to\n1/4 at the phase boundary between the Haldane phase\nand the FDC3 phase for small δB. At this phase bound-\nary, no finite size clusters coexist with cluster-3. There-\nfore, no pseudospin degrees of freedom can be defined.\nConsequently, no partial ferrimagnetic phase can be re-\nalized. In contrast,for largervalues of δB, we numerically\nfind a partialferrimagnetic phasebetween the FDC3 and\nUHphases. This wouldbe ascribed tothe contributionof\nother finite-length clusters with low lying energies which\ncome into play through higher-order processes in δB.\n5. Summary and Discussion\nWeintroducedtwotypesofdistortion,typeAandtype\nB, into the MDC with spins 1 and 1/2, and investigated\nthe ground-state phases. The phase diagrams are charac-\nteristicofthetypeAandtypeBdistortions,respectively.\nFor the type A distortion, the effective interaction be-\ntween the cluster spins is antiferromagnetic with bilinear\nand biquadratic terms. The numerically estimated val-\nues of the effective couplings show that the DC nground\nstatesaretransformedintotheHDC ngroundstates.The\norder parameters characterizing the HDC nphases are\ndefined and the UH-HDC nphase boundaries are deter-\nmined using the DMRG data. From the valence bond\nstructure of each phase, we expect that the UH-HDC n\nphase transition belongs to the universality class of the\n2-dimensional ( n+1)-clock model. The finite size scaling\nplot of the order parameter is consistent with this iden-\ntification. For the type B distortion, the effective inter-\naction between the cluster spins is ferromagnetic. In ad-\ndition to the FDC nphases with quantized spontaneous\nmagnetization m= 1/(n+1), the partial ferrimagnetic\nphasesarealsofoundnumericallybetweentheFDC nand\nFDC(n+1) phases. A physical interpretation of the par-\ntial ferrimagnetic phase is given for small δBby mapping\nonto an effective pseudospin-1/2 XXZ chain.\nGenerally, the introduction of lattice distortion into\na physical model increases the possibility that a corre-\nsponding material is realized. In the MDC, there are\nthree types of distortion modes affecting the exchange\ninteractions. Among them, the two types investigated in\nthe present paper are of generic nature, because the lo-\ncal conservation laws that hold in the undistorted MDC\nare broken. This suggests that the observation of the ex-\notic phenomena predicted in the present paper is pos-\nsible even if the corresponding material is not exactly\ndescribed by the model Hamiltonians (1) and (2).\nIf a distorted MDC material is synthesized, the dis-\ntortion may be controlled by, e.g., applying pressure. If\nthe distortion is of type A, the Curie constant vanishes\nas the DC nground state turns into one of the HDC n\nground states. The magnetic susceptibility and magnetic\nspecific heat will have an activation-type temperature\ndependence with activation energy proportional to the\neffective coupling between the cluster spins, which is of\nthe order of δA2. These HDC nphases are not realized\nif the distortion δAexceeds∼0.03 even in the most ro-\nbust case of n= 1. In a real material, the STSB in thevalence bond structure manifests itself as a magnetic su-\nperstructure. It is also possible that it is accompanied\nby a lattice superstructure of corresponding periodicity\nif the spin-lattice coupling is present. Therefore careful\nmeasurements of magnetic and lattice superstructures\nwould help with the observation of HDC nphases with\n1≤n≤3.\nOn the other hand, if the distortion of the material\nis of type B, the ground sate is ferrimagnetic. At low\nbut finite temperatures, however, the spontaneous mag-\nnetization vanishes owing to the one-dimensionality. As\na precursor of ordering at T= 0, the low-temperature\nmagnetic susceptibility should diverge as T−2with a co-\nefficient proportional to the effective coupling ∼δB2be-\ntween the cluster spins.52–54)This means that even a\nweak magnetic field of the order of H∼T2/δB2derives\nthe finite-temperature magnetization up to the value of\nthe ground-state spontaneous magnetization. This en-\nables the experimental estimation of the spontaneous\nmagnetization in real materials. The quantized ferrimag-\nnetic behavior should be observed for wide ranges of the\nparameters λandδBas shown in Fig. 11, and should\nbe easily observed if an appropriate material is synthe-\nsized. The partial ferrimagnetic phases are limited to\nnarrow intervals of the parameters δBandλ. Therefore,\nthesecanonlybeobservedasatemperature-independent\ncrossoverbetween two quantized ferrimagnetic behaviors\nwith careful exclusion of the thermal effect.\nWe have demonstrated that various exotic ground\nstates and phase transitions between them are realized\nin the distorted MDC with spins 1 and 1/2, which has\na strong frustration. The physical pictures of these phe-\nnomenahavebecomeclear.Thisismadepossiblebecause\nthe ground state of the undistorted MDC is known ex-\nactly. Therefore, we expect that our model may provide\na means of understanding the similar exotic phenomena\nrealized owing to the interplay of spin ordering,quantum\nfluctuation, and strong frustration in more general frus-\ntrated quantum chains on a firm ground. For example,\npartial ferrimagnetic phases are found in various one-\ndimensional frustrated quantum spin models.45–51)How-\never, some of them are only numerically confirmed and\nno physical explanation has been given so far. We hope\nthat the present study paves the way to the general un-\nderstanding of these partial ferrimagnetic states.\nWe thank J. Richter for drawing our attention to ref.\n28andrelatedworks.Thenumericaldiagonalizationpro-\ngram is based on the package TITPACK ver.2 coded by\nH. Nishimori. The numerical computation in this work\nhas been carried out using the facilities of the Super-\ncomputer Center, Institute for Solid State Physics, Uni-\nversity of Tokyo and Supercomputing Division, Informa-\ntion Technology Center, University of Tokyo. KH is sup-\nported by a Grant-in-Aid for Scientific Research on Pri-\nority Areas, ”Novel States of Matter Induced by Frustra-\ntion” (20048003) from the Ministry of Education, Cul-\nture, Sports, Science and Technology of Japan and a\nGrant-in-Aidfor Scientific Research(C) (21540379)from\nthe Japan Society for the Promotion of Science. KT and\nHS are supported by a Fund for Project Research from\nToyota Technological Institute.J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki 9\nAppendix\nThe Hamiltonian HBwith the type B distortion is\nrewritten as\nHB=H0+δH, (A·1)\nwhere\nδH=δBN/summationdisplay\nl=1/parenleftbig\nSl+Sl+1/parenrightbig/parenleftbig\nτ(1)\nl−τ(2)\nl/parenrightbig\n.(A·2)\nFor small δB, the ground state around λ=λc(n,n+1)\nconsists almost entirely of cluster- n’s and cluster-( n+\n1)’s. Hence, as a good approximation, we consider HB\nonly in the restricted Hilbert space where each state\ninvolves no clusters except for cluster- n’s and cluster-\n(n+ 1)’s. Under the fixed cluster number Ncin this\nHilbert space, H0is equivalent to the following effective\nHamiltonian expressed in terms of pseudospin operators:\nH0\neff=E0\nG(n+1;λ)Nc/summationdisplay\ni=1/parenleftbigg1\n2−σz\ni/parenrightbigg\n+E0\nG(n;λ)Nc/summationdisplay\ni=1/parenleftbigg1\n2+σz\ni/parenrightbigg\n, (A·3)\nwhereσz\ni= 1/2 andσz\ni=−1/2 correspond to ni=n\nandni=n+ 1, respectively. E0\nG(n;λ) is the ground-\nstate energy of a cluster- nand a dimer in the absence of\ndistortion, and is given by\nE0\nG(n;λ) =˜E(2n+1)+λn\n4−3λ\n4,(A·4)\nwhere˜E(2n+1) is the ground-state energy of the spin-1\nantiferromagneticHeisenbergchainwith length2 n+1.12)\nTheapplicationof δHtotheunperturbed groundstate\ntransforms one of the Tl= 0 bonds to a Tl= 1 bond or\nviceversa.Thentheresultingstatescontainclusterswith\nlengths less than nor greater than 2 n. Since these states\nare outside the restricted Hilbert space, no correction to\nthe ground-state energy is present within the first order\ninδB. Hence, the lowest-order correction is of the order\nofδB2. Up to the second order in δB, the effective pseu-\ndospin Hamiltonian is given by\nHeff=EG(n+1;λ,δB)Nc/summationdisplay\ni=1/parenleftbigg1\n2−σz\ni/parenrightbigg\n+EG(n;λ,δB)Nc/summationdisplay\ni=1/parenleftbigg1\n2+σz\ni/parenrightbigg\n+HXXZ,(A·5)\nwhereEG(n;λ,δB) is the ground-state energy of a\ncluster-nand a dimer including the second order cor-\nrection in δB. This is simply expressed as\nHeff=Nc¯EG+∆EGNc/summationdisplay\ni=1σz\ni+HXXZ,(A·6)\nwith\n¯EG=1\n2(EG(n+1;λ,δB)+EG(n;λ,δB)),(A·7)\n∆EG=EG(n;λ,δB)−EG(n+1;λ,δB).(A·8)The effective coupling constants Jeff\nzandJeff\n⊥inHXXZ\nare also of the second order in δB. We determine Jeff\nzand\nJeff\n⊥soastoreproducethelow-lyingenergyspectrumofa\npairofcluster- n’sbythatoftwo-pseudospinHamiltonian\nHeff(i,i+1) = 2 ¯EG+∆EG(σz\ni+σz\ni+1)+HXXZ(i,i+1)\n(A·9)\nIn each of the subspaces σz\ni+σz\ni+1=±1,σz\ni=\nσz\ni+1=σ=±1/2. Therefore, the Hilbert space is one-\ndimensional and the eigenvalue of Heff(i,i+ 1) is sim-\nplyEσσ= 2¯EG+ 2∆EGσ+Jeff\nz/4 withσ=±1/2.\nIn the subspace σz\ni+σz\ni+1= 0, the Hilbert space is\ntwo-dimensional and the eigenvalues of Heff(i,i+1) are\nE±= 2¯EG−Jeff\nz/4±Jeff\n⊥/2.\nThe original Hamiltonian of the cluster consisting of a\ncluster-nand a cluster- n′is the distorted diamond chain\nwith length n+n′.\nH(n+n′) =n+n′+1/summationdisplay\nl=1/bracketleftBig\n(1+δB)Slτ(1)\nl+(1+δB)τ(1)\nlSl+1\n+(1−δB)Slτ(2)\nl+(1−δB)τ(2)\nlSl+1+λτ(1)\nlτ(2)\nl/bracketrightBig\n,\n(A·10)\nWe denote the α-th eigenvalue of H(n+n′) asE(n+\nn′;α). Comparing the corresponding expression for the\neigenvalues, we find\nE(2n;0) =E1\n2,1\n2= 2¯EG+∆EG+Jeff\nz\n4(A·11)\nE(2n+2;0) = E−1\n2,−1\n2= 2¯EG−∆EG+Jeff\nz\n4(A·12)\nE(2n+1;0) = E−= 2¯EG−Jeff\nz\n4−Jeff\n⊥\n2(A·13)\nE(2n+1;1) = E+= 2¯EG−Jeff\nz\n4+Jeff\n⊥\n2.(A·14)\nSolving these sets of equations, with respect to Jeff\nzand\nJeff\n⊥, we find\nJeff\n⊥=E(2n+1;1)−E(2n+1;0), (A·15)\nJeff\nz= 2[E(2n+2;0)+ E(2n;0)\n−E(2n+1;1)−E(2n+1;0)].(A·16)\nNote that the rhs’s of (A ·15) and (A ·16) vanish for\nδB= 0. We numerically evaluated E(2n;0),E(2n+1;0),\nE(2n+1;1), and E(2n+2;0) at λ=λc(n,n+1) (n=\n0,1,2) for small δB. Using these values in eqs. (A ·15) and\n(A·16), we determined Jeff\n⊥andJeff\nzas eqs. (13)-(15).\nForthewholeMDC,theground-stateenergyiswritten\nas\nE0=Nc¯EG+Nc∆EGσ+NcǫXXZ(σ) (A ·17)\nwhereǫXXZ(σ) is the ground-state energy per site of a\nmagnetized spin-1/2 XXZ chain with /an}bracketle{tσz\ni/an}bracketri}ht=σ. The\nnumber of unit cells, N, of the original MDC is given by\nthe expectation value of eq. (9) as N=Nc(n+3\n2−σ).\nTherefore, we have\nE0=N\nn+3\n2−σ/parenleftbig¯EG+∆EGσ+ǫXXZ(σ)/parenrightbig\n.(A·18)10 J. Phys. Soc. Jpn. Full Paper K.Hida, K.Takano , and H. Suzuki\n−4 −2 0 2−0.500.5\n∆λ/Jeff\n⊥σn=0n=1\nn=2\nFig. A·1. Relationship between σand ∆λforn= 0,1, and 2.\nMinimizing this with respect to σwith fixed N, we find\n∆λ=/parenleftbigg\nn+3\n2−σ/parenrightbigg∂ǫXXZ(σ)\n∂σ+ǫXXZ(σ),(A·19)\nwhere ∆ λ=λ−λc(n,n+ 1;δB) andλc(n,n+ 1;δB) is\ndefined by\n(n+2)EG(n;λc,δB)−(n+1)EG(n+1;λc,δB) = 0.\n(A·20)\nTo simplify the calculation, we replace ǫXXZ(σ) by the\nground-state energy of the spin-1/2 XY chain ǫXY=\n−(Jeff\n⊥/π)cosπσ, because |Jeff\n⊥|is substantially larger\nthan|Jeff\nz|in all cases. Then we find\n∆λ\nJeff\n⊥=/parenleftbigg\nn+3\n2−σ/parenrightbigg\nsinπσ−1\nπcosπσ. (A·21)\nThis relation is plotted in Fig. A ·1 forn= 0,1 and 2. It\nis clear that σcontinuously increases from −1/2 to 1/2\nwith an increase in λ.\n1)Frustrated Spin Systems , ,ed. H. T. Diep: (World Scientific,\nSingapore, 2005), Chaps. 5 and 6.\n2)Proc. Int. Conf. on Highly Frustrated Magnetism (HFM2008)\nJ. Phys.: Conf. Series 145(2009).\n3) C. K. Majumdar and D. K. Ghosh: J. Math. Phys. 10(1969)\n1399.\n4) For examples of experimental materials, see M. Hase, H.\nKuroe, K. Ozawa, O. Suzuki, H. Kitazawa, G. Kido, and T.\nSekine: Phys. Rev. B 70(2004) 104426.\n5) B. S. Shastry and B. Sutherland: Physica B+C 108(1981)\n1069.\n6) H. Kageyama, K. Yoshimura, R. Stern, N.V. Mushnikov, K.\nOnizuka, M. Kato, K. Kosuge, C.P. Slichter, T. Goto, and Y.\nUeda: Phys. Rev. Lett. 82(1999) 3168.\n7) H. Kageyama, M. Nishi, N. Aso, K. Onizuka, T. Yosihama,\nK. Nukui, K. Kodama, K. Kakurai, and Y. Ueda: Phys. Rev.\nLett.84(2000) 5876.\n8) K. Takano: J. Phys. A: Math. Gen. 27(1994) L269.\n9) K. Takano, K. Kubo, and H. Sakamoto: J. Phys.: Condens.\nMatter8(1996) 6405.\n10) H. Niggemann, G. Uimin, and J. Zittartz: J. Phys.: Conden s.\nMatter9(1997) 9031.\n11) H. Niggemann, G. Uimin, and J. Zittartz: J. Phys.: Conden s.\nMatter10(1998) 5217.\n12) K. Takano, H. Suzuki, and K. Hida: Phys. Rev. B 80(2009)\n104410.\n13) K.Hida,K.Takano, and H.Suzuki: J.Phys.Soc.Jpn. 78(2009)084716\n14) K.Hida,K.Takano, and H.Suzuki:J.Phys.Soc.Jpn. 79(2010)\n044702.\n15) K. Okamoto, T. Tonegawa, Y. Takahashi, and M. Kaburagi:\nJ. Phys.: Condens. Matter 11(1999) 10485.\n16) K. Okamoto, T. Tonegawa, and M. Kaburagi: J. Phys.: Con-\ndens. Matter 15(2003) 5979.\n17) K. Sano and K. Takano: J. Phys. Soc. Jpn. 69(2000) 2710.\n18) H. Kikuchi, Y. Fujii, M. Chiba, S. Mitsudo, T. Idehara, T.\nTonegawa, K. Okamoto, T. Sakai, T. Kuwai, and H. Ohta:\nPhys. Rev. Lett. 94(2005) 227201.\n19) H. Ohta, S. Okubo, T. Kamikawa, T. Kunimoto, Y. Inagaki,\nH.Kikuchi, T.Saito, M.Azuma, and M.Takano: J.Phys.Soc.\nJpn.72(2003) 2464.\n20) A. Izuoka, M. Fukada, R. Kumai, M. Itakura, S. Hikami, and\nT. Sugawara: J. Am. Chem. Soc. 116(1994) 2609.\n21) D.Uematsu and M.Sato: J.Phys.Soc.Jpn. 76(2007) 084712.\n22) N. B. Ivanov, J. Richter, and J. Schulenburg: Phys. Rev. B 79\n(2009) 104412.\n23) N.B. Ivanov and J. Richter: Phys. Lett. A 232(1997) 308.\n24) J.Richter, N.B.Ivanov,and J.Schulenburg: J.Phys.:Co ndens.\nMatter10(1998) 3635.\n25) A. Koga, K. Okunishi, and N. Kawakami: Phys. Rev. B 62\n(2000) 5558.\n26) A. Koga and N. Kawakami: Phys. Rev. B 65(2002) 214415.\n27) J.Schulenburg and J.Richter: Phys.Rev.B 65(2002) 054420.\n28) J.Schulenburg and J.Richter: Phys.Rev.B 66(2002) 134419.\n29) T. Hakobyan, J. H. Hetherington, and M. Roger: Phys. Rev.\nB63(2001) 144433.\n30) L.˘Canov` a, J.Stre˘ cka, and M.Jas˘ c˘ ur: J.Phys.: Condens.Ma t-\nter18(2006) 4967.\n31) L.˘Canov` a, J. Stre˘ cka, and T. Lu˘ civjansk´ y: Condens. Matte r\nPhys.12(2009) 353.\n32) H.Kobayashi, Y.Fukumoto, and A.Oguchi: J.Phys.Soc.Jp n.\n78(2009) 074004.\n33) C. Mathoni` ere, J.-P.Sutter, and J. V.Yakhmi: in Magnetism:\nMolecules to Materials IV, ed. J. S. Miller and M. Drillon\n(Wiley, Weinheim, 2003) p. 1.\n34) Y. Hosokoshi and K. Inoue: in Carbon Based Magnetism , ed.\nT. L. Makarova and F. Palacio (Elsevier B. V., Amsterdam,\n2006) p. 107.\n35) E. Lieb and D. Mattis: J. Math. Phys. 3, (1962) 749.\n36) T. Kennedy: J. Phys.: Condens. Matter 2(1990) 5737.\n37) G. F´ ath and J. S´ olyom: Phys. Rev. B 44(1991) 11836.\n38) F. Y. Wu: Rev. Mod. Phys. 54(1982) 235.\n39) M. N. Barber: Phase Transitions and Critical Phenomena 8 ,\ned. C. Domb and J. L. Lebowitz (Academic Press, London,\n1983) p. 146.\n40) L.A. Takhtajan: Phys. Lett. 87A(1982) 479.\n41) H. M. Babujian: Phys. Lett. 90A(1982) 479.\n42) I. Affleck and F. D. M. Haldane: Phys. Rev. B36(1987) 5291.\n43) Y. Kato and A. Tanaka: J. Phys. Soc. Jpn 66(1997) 3944.\n44) A. Kitazawa and K. Nomura: J. Phys. Soc. Jpn. 66(1997)\n3944.\n45) S. Sachdev and T. Senthil: Ann. Phys. 251(1996) 76.\n46) L.Bartosch, M.Kollar,and P.Kopietz: Phys.Rev.B 67(2003)\n092403.\n47) N. B. Ivanov and J. Richter: Phys. Rev. B 69(2004) 214420.\n48) S. Yoshikawa and S. Miyashita: J. Phys. Soc. Jpn. Suppl. 74\n(2005) 71.\n49) K. Hida: J. Phys. Condens. Matter: 19(2007) 145225.\n50) K. Hida and K. Takano: Phys. Rev. B 78(2008) 064407.\n51) R.R.Montenegro-Filho and M.D.Coutinho-Filho: Phys.R ev.\nB78(2008) 014418.\n52) M. Takahashi: Prog. Theor. Phys. Suppl. 87(1986) 233.\n53) S. Yamamoto: Phys. Rev. B 59(1999) 1024.\n54) S. Yamamoto and T. Fukui: Phys. Rev. B 57(1998) R14008." }, { "title": "1303.1372v1.Change_in_the_Magnetic_Domain_Alignment_Process_at_the_Onset_of_a_Frustrated_Magnetic_State_in_Ferrimagnetic_La2Ni_Ni1_3Sb2_3_O6_Double_Perovskite.pdf", "content": "10-56 \n 1\nChange in the Magnetic Domain Alignment Process at the Onset of a \nFrustrated Magnetic State in Ferrimagnetic La 2Ni(Ni 1/3 Sb 2/3 )O6 \nDouble Perovskite \n(Revised ..) \n \nDiego G. Franco 1,2 , Raúl E. Carbonio 1, and G. Nieva 2,3 \n \n1INFIQC-CONICET, Depto. de Físico Química, Facultad de Cienc ias Químicas, Universidad Nacional de Córdoba, Ciudad \nUniversitaria. X5000HUA Córdoba, Argentina \n2Laboratorio de Bajas Temperaturas. Centro Atómico Baril oche –CNEA. 8400 Bariloche, R. N., Argentina. \n3Instituto Balseiro, CNEA and Universidad Nacional de Cuy o. 8400 Bariloche, R. N., Argentina. \n \nWe have performed a combined study of magnetization hysteresis loops and time dependence of the magnet ization in a broad \ntemperature range for the ferrimagnetic La 2Ni(Ni 1/3 Sb 2/3 )O6 double perovskite. This material has a ferrimagne tic order transition \nat ~100 K and at lower temperatures ( ~ 20 K) shows the signature of a frustrated state du e to the presence of two competing \nmagnetic exchange interactions. The temperature dep endence of the coercive field shows an important up turn below the point \nwhere the frustrated state sets in. The use of the magnetization vs. applied magnetic field hysteresis data, together with the \nmagnetization vs. time data provides a unique oppor tunity to distinguish between different scenarios f or the low temperature \nregime. From our analysis, a strong domain wall pin ning results the best scenario for the low temperat ure regime. For \ntemperatures larger than 20K the adequate scenario seems to correspond to a weak domain wall pinning. \n \nIndex Terms — Ferrimagnetic materials, Magnetic analysis, Magneti c domain walls, Magnetic hysteresis. \n \nI. INTRODUCTION \nANY of the magnetic interactions found in transition \nmetal oxide perovskites are due to superexchange \nand/or super-superexchange interactions mediated through \nthe O2- p orbitals. In some materials the relative strength of \nthese interactions determines the magnetic structure, range \nof the ordering temperatures and the possibility of \nfrustration [1]-[5]. In the perovskite structure, the t ypical \nbond angles and distances usually favor antiferromagnetic \nsuperexchange interactions [6],[7]. However, in some \nspecial cases, due to disorder or differences in the magne tic \nstate of the cations, a ferrimagnetic state is develope d with \nmacroscopic characteristics similar to a ferromagnetic state. \nCoercivity and remanence are an indication of the \nmetastability in ferromagnetic samples. Their magnitudes \nindicate how far the system is from equilibrium. They ar e \nrelated, therefore, with the relaxation to the equilibriu m \nstate, the anhysteretic [8] curve in the ferromagnetic st ate. In \nbulk ferromagnets the energy barriers that determine the \ntime evolution of the magnetization are related to local \ninteractions within a domain, the nucleation and the \nmovement of domain walls (DW). The DW movement \ndepends on the applied magnetic force, wall thickness and \ntype and density of pinning centers. \nIn bulk ferromagnetic samples, a local frustration is \nnormally hard to visualize due to the magnetic history \ndependence of the metastable states. However, the magneti c \nmoments alignment within a domain and the movement of \nthe DW have characteristic energies [9] that could be \nmodified if some degree of magnetic frustration occurs at a microscopic level. This normally results in a strong DW \npinning effect and causes an increase in the coercivity. \nThis article will present a detailed magnetic study of t he \nferrimagnetic double perovskite La 2Ni(Ni 1/3Sb 2/3)O 6. We \nwill show that the material behaves as a ferrimagnet below \n100 K and that there is a change in the magnetic domains \nalignment process at 20 K. We will show that below 20 K \nthe hysteretic magnetic behavior is characteristic of a strong \ndomain wall pinning regime due to the onset of a frustrated \nmagnetic interaction. \nII. RESULTS \nWe prepared polycrystalline samples of \nLa 2Ni(Ni 1/3 Sb 2/3 )O6 by conventional solid-state reaction at \n1400 oC [10]. X ray diffraction data from powders at room \ntemperature showed the crystalline symmetry to be \nmonoclinic, space group P2 1/n. This space group \naccommodates a rock salt arrangement of BO 6 and B'O 6 \noctahedra described by the a -b-c+ system of three octahedral \ntilts in the Glazer's notation. The (Ni/Sb) 2dO6 and \n(Ni/Sb) 2cO6 octahedra are rotated in phase (along the \nprimitive c axis) and out-of phase (along the primitive a and \nb axes). We performed a Rietvelt refinement of the structur e \nusing the FULLPROF program [11], resulting in lattice \nparameters of a = 5.6051(3) Ǻ, b = 5.6362(3) Ǻ, c = \n7.9350(5) Ǻ and β = 89.986(4) o. We refined the two \ncrystallographic sites 2 d and 2 c with different occupancies \nNi 2+ /Sb 5+ to model the octahedral site disorder. The 2 d \ncation site is almost fully occupied by Ni 2+ whereas the 2 c \nsite has occupancy close to 1/3 of Ni 2+ ions and 2/3 of Sb 5+ . \nThe resulting crystallographic formula can be written as \nLa 2(Ni 0.976 Sb 0.024 )2d(Ni 0.357 Sb 0.643 )2cO6. M\nManuscript received January XX, 2013 (date on which paper was \nsubmitted for review). Corresponding author: G. Nie va (e-mail: \ngnieva@cab.cnea.gov.ar). \nDigital Object Identifier inserted by IEEE 10-56 \n 2\nThe magnetic measurements were performed on \npolycrystalline pellets with a QD-MPMS SQUID \nmagnetometer in the range 2 to 300K and -5 to 5T. In the \nmain panel of Fig. 1 we show the magnetization, M, as a \nfunction of temperature, T, while cooling in a very low \napplied field, H. There is a transition to a magnetic \npolarized state at TC = 98(2) K. \nWe extrated the low temperature value of the saturation \nmagnetization, Ms, from M vs. H curves, from the \nasymptotic extrapolation of the high field behavior with a \nLangevin function. This saturation magnetization, Ms, has a \nlower value than the one expected for the complete \npolarization of the Ni 2+ magnetic moments, 2.67 µB/f.u.. \nInstead, the experimental Ms value was 1.19 µB/f.u., \nimplying that the system behaves as a ferrimagnet, with two \nNi 2+ magnetic sublattices antiferromagnetically coupled, one \nat the 2 d site and another at the 2 c site. The near 1/3 Ni 2+ \nrandom occupation of the 2 c sites sublattice give as a result \nuncompensated Ni 2+ magnetic moments that order at 100 K. \nFor a perfectly stoichiometric ferrimagnetic sample and ful l \nNi 2+ occupancy of the 2 d site Ms should be 1.33 µB/f.u., and \nlower values are expected if Sb 5+ partially occupies also the \n2d site. The expected value for Ms with the refined \noccupancies is 1.24 µB/f.u., very close to the experimental \none. \nWe measured hysteresis loops, M vs H, for several \ntemperatures below 100 K. We show in the inset of Fig. 1 a \ndetail of the loops for 2 K and 20 K. \nFIG. 1 HERE \nWe have also measured the time evolution of the \nmagnetization at the coercive field (i.e. near the field f or \nzero magnetization) after saturation at 1, 3 and 5 T for each \ntemperature. We show in Fig. 2 typical M vs time data, for \nthree different temperatures at their corresponding coercive \nfields. \n \nFIG. 2 HERE \nIII. DISCUSION \nWe show in Fig. 3(a) and (b) the temperature dependence \nof the coercive field, Hc, and the ratio between remanent \nmagnetization and saturation magnetization, ( Mr / M s). The general feature observed in Fig. 3 is that Mr and Hc increase \nsteeply when the temperature is lowered below 20 K \nindicating an increase in the energy absorbed by the materi al \nto change the direction of M. \nThe measured values of the coercive field, Hc, display two \ndifferent regimes as can be seen in Fig. 3(a). For T > 20 K a \nlinear behavior of Hc was found. This linear behavior is \ncharacteristic of weak DW pinning (WDWP), produced by a \nrandom distribution of individual weak pinning sites [9]. In \nthis case the coercive field is given by \n \n\n\n\n\n\n\n\n− =2 031 25 1\nbTkH HB\nW cγ (1) \n \nwhere H0W is the zero temperature extrapolated reversion \nfield, kB is the Boltzmann constant, γ is the DW energy per \nunit area and b is a measure of the DW thickness. The \nobtained values are shown in Table I. \nIn the low temperature regime, T < 20 K, two models \ndescribe reasonably well the data. One corresponds to stron g \nDW pinning (SDWP), \n \n23 / 2\n0475 1\n\n\n\n\n\n\n\n− =bf TkH HB\nS c (2) \n \nwhere H0S is the coercive field at zero temperature and f is \nthe magnetic force needed to depin a domain wall. The \nfitted values are shown in Table I. \nThe other model corresponds to the freezing of single \ndomain large particles (SDLP) or clusters [11], [12]. In t his \nscenario, \n \n\n\n\n\n\n\n− =2/ 125 1KV TkH HB\nK c (3) \n \nwhere HK is the anisotropy field of a particle or cluster, V is \nits volume and K is the uniaxial anisotropy energy density. \nThe fitted values are shown in Table I. \nFIG. 3 HERE \n \n \nFig. 1. (color online) Magnetization as a function of temperature cooling \nwith an applied field of 1 Oe. Inset: Magnetizatio n as a function of ap plied \nfield, detail of the magnetization loops for two fi xed temperatures T = 2 K \nand 20 K. \n \nFig. 2. (color online) Difference between the measured magnetization and \nthe initial one, M 0, as a function of time. The shown mag netization time \ndependence was taken at H c (M ~ 0) after saturation in the opposite \ndirection. We indicate the fixed temperatures for each experiment. 10-56 \n 3\nIn Fig. 3(a) and the inset we show the lines (solid and \ndash doted) corresponding to each model. The best fit is \nobtained with the SDWP model but the freezing of SDLP \nmodel is also in fair agreement with the Hc data. \nThe time evolution of the magnetization could be used to \ndiscern between the two scenarios at low temperature. I f a \ndistribution of activation energies is present in the mater ial \n[9] a logarithmic behavior is expected for M(t): \n \n)ln( 0 t S M M − = (4) \n \nwhere M0 is the starting value of the magnetization and S the \nmagnetic viscosity coefficient. The above relation holds \napproximately for our polycrystalline pellets samples a s we \nshow in Fig. 2. \nIn a tipical ferromagnet, the time dependence of M is \nirreversible and this behavior has been connected with the \nirreversibility caused by a small change in field, the so \ncalled irreversible susceptibility, χirr . Both irreversibilities \nare related by a fictitious field, the fluctuation field, Hf, in \nthe theory introduced by Neel [13] that represents an \naverage of the thermally activated, time dependent processes \n[14] leading to equilibrium by reversing the metastable \nmagnetization. In terms of the magnetization derivatives, a t \na given field and temperature, \n \nf irr \nMi\nt iirr \nHirr HtH\nHM\ntMS\nirr iχ=\n\n\n\n∂∂\n\n\n\n\n∂∂=\n\n\n\n∂∂=)ln( )ln( (5) \n \nwhere Mirr is the irreversible magnetization and Hi is the \ninternal field. In the case of a time independent viscosity coeficient S, the fluctuation field is equivalent to the \nmagnetic viscosity parameter Sv, that can be written in terms \nof the activation energy, E, necessary for magnetization \nreversal [15], \n \nirr MB\nvS\ndH dE TkS\nirr χ=−=) /(. (6) \n \nTo determine the temperature dependence of Sv, a value of \nH equal to the coercive field is chosen [15] ( Mirr = 0). The \nactivation energy for the SDWP, WDWP and clusters or \nSDLP freezing are given by [16]: \n \n2/32/1\n01) 3 / 4 (\n\n\n\n\n\n\n\n− =\nSSHHbf E (7) \n\n\n\n\n\n\n\n− =\nWWHHb E\n021) 31 (γ (8) \n2\n1\n\n\n\n\n\n\n\n− =\nKFHHKV E (9) \n \nand in each case the magnetic viscosity parameters are g iven \nrespectively by, \n \n\n\n\n\n\n\n\n−\n\n\n=3/2 3/2\n0475 1475 \n75 4)(bf Tk\nbf TkH SB B\nS Sv (10) \n2 031 25 \n25 1)(bTkH SB\nW Wvγ= (11) \n2/125 \n50 1)( \n\n=KV TkH SB\nK Fv. (12) \n \nFrom the experimental data (such as those of the inset of \nFig. 4) the values of χirr (H = Hc) can be extracted. They can \nbe approximated as those of the total χ at Hc, neglecting the \nreversible contribution to χ [14]. These values are shown in \nFig. 4. Also from the experimental data the S values can be \ncalculated from (4) at Hc, since the linear behavior holds, \nFig. 2. In this case we take Mirr as the measured M, \nneglecting the reversible component. \nFIG. 4 HERE \nTABLE 1 HERE \nThe experimental values of Sv were obtained by using (6). \nThey are displayed in Fig. 5 together with the fits for \ndifferent models in different temperature ranges using (10) - \n(12). At low temperature, the best fit to the data is given \nTABLE I \n γb2 \n(10 -14 erg) H0W \n(Oe) 4bf \n(10 -13 erg) H0S \n(Oe) KV \n(10 -14 erg) H0F \n(Oe) \nHc 1.26 53.5 3.07 780 7.4 760 \nSv 1.4 53.5 2.13 780 7.4 760 \n \nFitted parameters for the WDWP model above 20 K and the two models \ncompared below 20K, SDWP and freezing of SDLP. The first column \nindicates whether the coercive field or the magneti c viscosity parameter \nwere used in the parameters determination. In the case of the viscosity \nparameter, the zero temperature fields H0W , H0S and H0F were not fitted but \ntaken from the Hc fits. \n \n \nFi g. 3. (color online) (a) Coercive field Hc vs T for La 2Ni(Ni 1/3Sb 2/3)O 6\npolycrystalline samples. The solid lines (SDWP mode l), dash- dotted lines \n(SDLP freezing model) at low temperature and the da shed line (WDWP \nmodel) at high temperature are fits described in th e text. Inset: H c 1/2 vs T 2/3 \nshowing the linear behavior expected in the SDWP mo del, solid lines are \nthe SDWP model and dash- dotted lines are the SDLP freezing model. (b) \nNormalized remanent magnetization (at H = 0) vs T. The different \nsymbols indicate different samples in both panels. 10-56 \n 4\nusing (10) (a SDWP scenario) provided the nonmonotonic \nbehavior of Sv. The parameters obtained are shown in Table \nI. Clearly no good agreement is found for the freezing of \nclusters or SDLP scenario. In the high temperature region \nthe experimental Sv vs T is in agreement with the linear \nbehavior calculated in (11). However, a non-zero Sv (T = 0) \nvalue was found, not present in the model. \nFIG. 5 HERE \nTherefore, based on combined data extracted from the \nhysteresis loops and time dependence of the magnetization, \nwe depicted two regimes in La 2Ni(Ni 1/3Sb 2/3)O 6 pelletized \npolycrystalline samples: Weak pinning of DWs at T >2 0 K \nand strong pinning of DWs below that temperature. The \nmicroscopic origin of this change of regime could be related \nwith the onset of a super-superexchange antiferromagnetic \ninteraction among Ni 2+ via O 2- -Sb 5+ - O2- paths [10] that \ncreates a frustrated magnetic interaction. IV. CONCLUSION \nThe ferrimagnetic state in La 2Ni(Ni 1/3Sb 2/3)O 6 was found \nto be characterized by two different regimes for domain w all \nmovement, a strong and a weak domain wall pinning regime \nat low and high temperatures respectively. The temperature \nrange of the strong domain wall pinning regime coincides \nwith that of the existence of a proposed frustrated stat e. The \nscenario of clusters or large single domain particles f reezing \nwas discarded based in the coercive field and magnetic \nviscosity parameter temperature dependence analysis. \nACKNOWLEDGMENT \nWe thank E.E. Kaul for fruitfull discussions. R.E.C, and \nG.N. are members of CONICET. D.G.F. has CONICET \nscholarship. Work partially supported by ANPCyT PICT07-\n819, CONICET PIP 11220090100448 and SeCTyP-\nUNCuyo 06/C313. R.E.C. thanks FONCYT (PICT2007-\n303), CONICET (PIP 11220090100995) and SECYT-UNC \n(Res. 214/10) for finantial support. \nREFERENCES \n[1] H.-J. Koo and M.-H. Whangbo, “Importance of the O-M -O bridges \n(M = V5+, Mo 6+) for the spin-exchange interactions in the magneti c \noxides of Cu 2+ ions bridged by MO 4 tetrahedra: Spin-lattice models of \nRb 2Cu 2(MoO 4)3,BaCu 2V2O8, and KBa 3Ca 4Cu 3V7O28 ,” Inorg. Chem. \n45, 4440(2006). \n[2] M. del C. Viola, M. S. Augsburger, R. M. Pinacca, J . C. Pedregosa, \nR.E. Carbonio and R. C. Mercader, “Order-disorder at F e sites in \nSrFe 2/3 B” 1/3 O3 (B”=Mo, W, Te, U) tetragonal double perovskites,” J. \nSolid State Chem. 175, pp. 252-257 (2003). \n[3] A. Maignan, B. Raveau, C. Martin and M. J. Hervieu, “L arge \nintragrain magnetoresitance above room temperature in the double \nperovskite Ba 2FeMoO 6,” J. Solid State Chem. 144, pp. 224-227 \n(1999). \n[4] S. H. Kim and P.D. Battle, “Structural and electron ic properties of the \nmixed Co/Ru perovskites AA’CoRuO 6 (A, A’=Sr, Ba, La),” J. Solid \nState Chem. 114, pp. 174-183 (1995). \n[5] P. S. R. Murthy, K. R. Priolkar, P. A. Bhobe, A. Das, P. R. Sarode \nand A. K. Niga, “Disorder induced negative magnetiz ation in \nLaSrCoRuO 6,” J. of Magn. and Magn. Mater . 322, pp. 3704-3709 \n(2010). \n[6] J. B. Goodenough. Magnetism and the Chemical Bond , Interscience, \nNew York. N.Y. 1963. \n[7] P. J. Hay, J. C. Thibeault and R. J. Hoffmann, “Orb ital interactions in \nmetal dimer complexes,” J. Am. Chem. Soc. 97, pp. 4884-4899 \n(1975). \n[8] D. C. Jiles and D. L. Atherton, “Theory of ferromagn etic hysteresis” \nJ. of Magn. and Magn. Mater. 61, pp. 48-60 (1986). \n[9] P. Gaunt, “Ferromagnetic domain wall pinning by a r andom array of \ninhomogeneities,” Phil. Mag. B 48, pp. 261-276 (1983). \n[10] D. G. Franco, G. Nieva and R. E. Carbonio, to be pub lished \nelsewhere. \n[11] X. -G. Li, X. J. Fan, G. Ji, W. B. Wu, K. H. Wong, C. L.Choy, and H. \nC. Ku, “Field induced crossover from cluster-glass t o ferromagnetic \nstate in La 0.7 Sr 0.3 Mn 0.7 Co 0.3 O3,” J. Appl. Phys. 85, pp. 1663-1666 \n(1999). \n[12] J. García-Otero, A. J. García-Bastida and J. Rivas, “ Influence of \ntemperature on the coercive field of non-interactin g fine magnetic \nparticles,” J. Magn. Magn. Mater. 189, pp. 377-383 (1998). \n[13] E. P. Wohlfarth, “The coefficient of magnetic visco sity,” J. Phys. F. \nMet. Phys. 14, pp. L155-L159 (1984). \n[14] O. V. Billoni, E. E. Bordone, S. E. Urreta, L. M. Fab ietti, H. R. \nBertorello, “Magnetic viscosity in a nanocrystalline two phase \ncomposite with enhanced remanece,” J. of Magn. Magn. Mater. 208, \npp. 1-12 (2000). \n[15] D. C. Crew, P.G. McCormick, R. Street, “Temperature D ependence \nof the Magnetic Viscosity Parameter,” J. of Magn. Magn. Mater. 177-\n181, pp. 987-988 (1998). \n[16] J. F. Liu and H. L. Luo, “On the coercive force and effective \nactivation volume in magnetic materials,” J. of Magn. Magn. Mater. \n94, pp. 43-48 (1991). \nFig. 5. (color online) Magnetic viscosity paramete r Sv vs T for \nLa 2Ni(Ni 1/3Sb 2/3 )O 6 polycrystalline pellets. Different symbols indicate \ndifferent samples, the solid symbols corresponds to Sv values cal culated \nwith the data displayed in Fig.2 . The lines represent the models described in \nthe text, SDWP model (solid line), freezing of SDLP model (dash dot ted \nline) and WDWP model (dash line). In the inset , a zoom of the low \ntemperature region is shown. \nFig. 4. (color online) Total susceptibility χ= dM/dH measured at Hc vs T \nfor the polycrystalline pellets. Different symbols indicate different samples. \nThe line is a guide to the eye. The inset shows a t ypical χ vs H curve taken \nat 10 K. " }, { "title": "1911.05270v3.Variety_of_order_by_disorder_phases_in_the_asymmetric__J_1_J_2__zigzag_ladder__From_the_delta_chain_to_the__J_1_J_2__chain.pdf", "content": "Variety of order-by-disorder phases in the asymmetric J1\u0000J2zigzag ladder: From the delta chain\nto the J1\u0000J2chain\nTomoki Yamaguchi,1Stefan-Ludwig Drechsler,2Yukinori Ohta,1and Satoshi Nishimoto2, 3\n1Department of physics, Chiba University, Japan\n2Institute for Theoretical Solid State Physics, IFW Dresden, 01069 Dresden, Germany\n3Department of Physics, Technical University Dresden, 01069 Dresden, Germany\n(Dated: March 13, 2020)\nWe study an asymmetric J1-J2zigzag ladder consisting of two different spin-1\n2antiferromagnetic (AFM; J2,\n\rJ2>0) Heisenberg legs coupled by zigzag-shaped ferromagnetic (FM; J1<0) inter-leg interaction. On the\nbasis of density-matrix renormalization group based calculations the ground-state phase diagram is obtained as\nfunctions of \randJ2=jJ1j. It contains four kinds of frustration-induced ordered phases except a trivial FM\nphase. Two of the ordered phases are valence bond solid (VBS) with spin-singlet dimerization, which is a rather\nconventional order by disorder. Still, it is interesting to note that the VBS states possess an Affleck-Kennedy-\nLieb-Tasaki-type topological hidden order. The remaining two phases are ferrimagnetic orders, each of which\nis distinguished by commensurate or incommensurate spin-spin correlation. It is striking that the ferrimagnetic\norders are not associated with geometrical symmetry breaking; instead, the global spin-rotation symmetry is\nbroken. In other words, the system lowers its energy via the FM inter-leg interaction by polarizing both of the\nAFM Heisenberg legs. This is a rare type of order by disorder. Besides, the incommensurate ferrimagnetic state\nappears as a consequence of the competition between a polarization and a critical Tomonaga-Luttinger-liquid\nbehavior in the AFM Heisenberg legs.\nI. INTRODUCTION\nLow-dimensional frustrated quantum magnets, in which a\nmacroscopic number of quasi-degenerate states compete with\neach other, provide an ideal playground for the emergence of\nexotic phenomena1. For instance, the interplay of frustration\nand fluctuations could lead to unexpected condensed matter\norders at low temperatures by spontaneously breaking some\nsort of symmetry order by disorder2. Long-range ordered\n(LRO) magnetic state with breaking a spatial symmetry as\nwell as valence bond solid (VBS) state with a formation of\ndisentangled-unit–like local spin-singlet pair are typical ex-\namples of order by disorder. Moreover, when quantum fluctu-\nations between the quasi-degenerate states prevent a selection\nof particular order, one ends up with spin liquids. Modern the-\nories have brought us new insight by identifying spin liquids\nas topological phases of matter3,4. In recent years the realiza-\ntion of topological phases in frustrated spin systems has been\none of the central topics in condensed matter physics5–7.\nIn one-dimensional (1D) and spin-1\n2case, quantum fluc-\ntuations are maximized so that we may place more expec-\ntations on the discovery of novel ground states by the co-\noperative effects with magnetic frustration. A most simply-\nstructured 1D frustrated system is the so-called spin-1\n2J1-\nJ2chain consisting of nearest-neighbor J1and next-nearest-\nneighborJ2couplings. When both J1andJ2are antiferro-\nmagnetic (AFM), the ground state is a VBS, the nature of\nwhich can be grasped by the Majumdar-Ghosh (MG) model8,\natJ2=J1>\u00180:249,10. The idea of MG model was generalized\nto the Affleck-Kennedy-Lieb-Tasaki (AKLT) model11exhibit-\ning spin- 1VBS ground state with a symmetry protected topo-\nlogical order12.\nMeanwhile, the J1-J2chain with ferromagnetic (FM)\nJ1and AFMJ2, which is known as a standard magnetic\nmodel for quasi-1D edge-shared cuprates such as Li 2CuO 213,\nLiCuSbO 414, LiCuVO 415, Li2ZrCuO 216, Rb 2Cu2Mo3O1217,and PbCuSO 4(OH) 218, encloses wider array of states of mat-\nter. Theoretically, this model has been extensively studied:\nOther than a trivial FM state in the dominant J1region, the\nground state is a topological VBS accompanied by sponta-\nneous multiple dimerization orders19,20(More details are de-\nscribed in Sec. III A of this article). It is also intriguing that\na vector chirality and multimagnon bound states are induced\nin the presence of magnetic field21–24. Especially, the detec-\ntion of nematic or higher multipolar phases is one of the most\nexciting experimental current issues14,25–29. Sensitive features\nto even tiny interchain couplings are another characteristic of\nthis system30–32.\nAnother typical example of frustrated 1D system is\nthe delta chain (or sawtooth chain). The lattice structure\nis a series of triangles, as shown in Fig. 1(b), which is\nsimilar to that of the J1-J2chain but certain parts of\nJ2bonds are missing. There have been several candi-\ndates for delta-chain and related materials: YCuO 0:2533,34,\n[Cu(bpy)(H2O)][Cu( bpy)(mal)(H2O)](ClO 4)235, Zn L2S4\n(L= Er, Tm, and Yb)36, Cu(AsO 4)(OH)\u00013H2O37,\nMn2GeO 438, Rb 2Fe2O(AsO 4)239, CuFe 2Ge240, Fe10Gd1041,\nCu2Cl(OH) 342, Fe2O(SeO 3)243, and V 6O1344. In these mate-\nrials, a wide array of complex phases has been experimentally\nobserved. Delta-chain systems may offer an outlook towards\npromising prospects on novel magnetic phenomena.\nThe magnetic properties of delta-chain systems are totally\ndifferent in different signs of J1andJ2. For the case of\nJ1<0andJ2>0(typically, referred as FM-AFM delta\nchain), only two corresponding materials have been recog-\nnized. One of them is malonatobridged copper complexes\n[Cu(bpy)(H2O)][Cu( bpy)(mal)(H2O)](ClO 4)235. The mag-\nnetic Cu2+ions with effective spin-1\n2form into a delta-chain\nnetwork. The base and the other exchange couplings in a tri-\nangle made of malonate were estimated, respectively, as AFM\n(J2= 6:0K) and FM ( J1=\u00006:6K) from the analysis of mag-\nnetic susceptibility \u001f(T); and, as AFM ( J2= 10:9K) and FMarXiv:1911.05270v3 [cond-mat.str-el] 12 Mar 20202\n(J1=\u000012:0K) from the fitting of magnetization curve (typ-\nically, referred as FM-AFM delta chain). In either case, the\nratio of AFM and FM couplings is close to 1. This means that\nthe material would be in the region of strong frustration. The-\noretically, the ground state was predicted to be a ferrimagnetic\nstate but the detailed spin structures are less understood45.\nIn fact, only qualitative behavior of measured magnetization\ncurve could be explained by assuming a ferrimagnetic ground\nstate46,47. A deeper understanding of the ferrimagnetic state\nis necessary to resolve the remaining discrepancy between ex-\nperiment and theory.\nThe second candidate of the FM-AFM delta-chain ma-\nterials is a mixed 3 d=4fcyclic coordination cluster system\nFe10Gd1041. In these days, the delta-chain physics is increas-\ningly attracting attention due to the synthesis of Fe 10Gd10.\nThis cluster consists of 10 + 10 alternating Gd and Fe ions.\nThe exchange couplings were estimated as FM ( J1=\u00001:0K)\nbetween Fe and Gd ions, AFM ( J2= 0:65K) between Fe\nions, and nearly zero between Gd ions; the magnetic ions\nform an FM-AFM delta chain short ring. The parameter ratio\nJ2=jJ1j= 0:65seems to be very close the FM quantum criti-\ncal pointJ2=jJ1j= 0:748Although the spin values of Fe and\nGd ions are higher than S= 1=2(S= 5=2andS= 7=2, re-\nspectively), quantum fluctuations would play important roles\nto determine the ground state because of the quantum critical-\nity49. This means that the magnetic properties can be drasti-\ncally changed upon even a small variation of external influ-\nences such as magnetic field, pressure, chemical means, and\ngating current. So, this delta chain material is drawing atten-\ntion also from the perspective of controlling magnetic states\nin molecular spintronics50.\nFor comparison, a few examples of delta-chain materials\nonly with AFM interactions ( J1;J2>0) have been also\nreported. With the help of MG-like projection method, the\nmagnetic properties of the AFM-AFM delta-chain are better\nunderstood than those of the FM-AFM one33,51–53. A pecu-\nliarly interesting feature is the dispersionless kink-antikink\ndomain wall excitations to the dimerized VBS ground state.\nA kink is highly localized only in the range of one triangle.\nThe first candidate of AFM-AFM delta-chain materials was\nthe delafossite YCuO 2:533. However, a first-principle calcula-\ntion revealed that the ratio of J2=J1in YCuO 2:5is out of the\nrange of the dimerized VBS ground state and additional in-\ntrachain FM interaction is significantly large34. Very recently,\nthe other candidate materials Cu 2Cl(OH) 3(S= 1=2)42and\nFe2O(SeO 3)2(S= 5=2) have been reported. They indeed\nexhibit characteristic features of AFM-AFM delta chain: a\nmagnetization plateau at half-saturation46,54in Cu 2Cl(OH) 3\nand an almost flat-band one-magnon excitation spectrum in\nFe2O(SeO 3)2.\nAs mentioned above, the research of frustrated 1D systems\nwithJ1-J2or delta-chain structures has become more and\nmore active. Interestingly, each of the J1-J2chain and the\ndelta chain is expressed as a limiting case of an asymmetric\nJ1-J2zigzag ladder, defined as two different AFM Heisen-\nberg chains coupled by zigzag-shaped interchain FM interac-\ntion [see Fig. 1(a)]. When one of the Heisenberg chains van-\nishes, it is the delta chain; and, when the Heisenberg chainsare equivalent, it is the J1-J2chain. However, it is known that\ntheir ground states are completely different. Then, one may\nsimply question how the two limiting cases are connected.\nOf particular interest is that the effect of exchange coupling\ntowards the J1-J2chain can be a likely perturbation in real\ndelta-chain compounds, e.g., the effect of tiny coupling be-\ntween Gd ions in Fe 10Gd10.\nIn this paper, we therefore study an asymmetric FM-AFM\nJ1-J2zigzag ladder using the density-matrix renormalization\ngroup (DMRG) technique. We first clarify the detailed spin\nstructure and low-energy excitations of ferrimagnetic state in\nthe delta chain limit. We suggest that the ferrimagnetic state is\na rare type of order by disorder, where the energy is lowered\nby FM fluctuation between two polarized AFM Heisenberg\nchains with spontaneous breaking of the global spin-rotation\nsymmetry. Then, we examine how the ferrimagnetic state is\ncollapsed and connected to the well-known incommensurate\nspiral state in the J1-J2chain. We also find there exist two\nkinds of VBS phases in the spiral region. Finally, we obtain\nthe ground-state phase diagram of asymmetric J1-J2zigzag\nladder with interpolating between the delta chain and J1-J2\nchain.\nThe paper is organized as follows: In Sec. II our spin\nmodel is explained and the applied numerical methods are de-\nscribed. In Sec. III we briefly mention to-date knowledge on\nthe ground state for two limiting cases of our spin model. In\nSec. IV we present our numerical results and discuss how the\ntwo limiting cases are connected. Finally we end the paper\nwith a summary in Sec. V .\n(a)\n(b)A\nB\nA\nB\n(c)\nA\nB\nFIG. 1. (a) Lattice structure of the asymmetric J1-J2zigzag ladder.\nThe indices ‘ A’ and ‘ B’ denote apical and basal chains, respectively.\nThe lattice spacing ais set as a distance between neighboring sites\nalong the chains. The AFM interaction in the apical chain is con-\ntrolled by\r. (b) Lattice structure of the so-called delta chain (or saw-\ntooth chain) which is realized in the limit of \r= 0. (c) Schematic\nrepresentation of the ferrimagnetic state with global spin-rotation-\nsymmetry breaking.3\nII. MODEL AND METHOD\nA. Model\nThe asymmetric J1-J2zigzag ladder is defined as two\nHeisenberg chains coupled by zigzag-shaped interchain inter-\naction. The lattice structure is sketched in Fig. 1(a). We call\na leg chain with larger interaction “basal chain” and the other\nwith smaller interaction “apical chain”. The Hamiltonian is\nwritten as\nH=J1X\niSA;i\u0001(SB;i+SB;i+1)\n+J2X\ni(SB;i\u0001SB;i+1+\rSA;i\u0001SA;i+1); (1)\nwhere SB;iis spin-1\n2operator at site ion the basal chain and\nSA;iis that on the apical chain. We focus on the case of\nFM interchain coupling ( J1<0) and AFM intrachain cou-\npling (J2>0). The intrachain interaction of apical chain is\ncontrolled by \r(0\u0014\r\u00141). The system (1) corresponds\nto the so-called delta chain (or sawtooth chain) at \r= 0\n[Fig. 1(b)] and the so-called J1-J2chain at\r= 1. The exist-\ning knowledge on the ground-state properties of these chains\nis briefly summarized in the next section. In our numerical\ncalculations the chain lengths of basal and apical chains are\ndenoted asLBandLA, respectively. The total number of\nsites isL=LB+LA. In this paper, we call the system\nfor0< \r < 1“asymmetric J1-J2zigzag ladder”, which has\nbeen little or not studied. In the case of J1>0andJ2>0,\nthere are a few studies55,56.\nB. DMRG methods\nIn order to examine the ground state and low-energy\nexcitations of asymmetric J1-J2zigzag ladder, we em-\nploy the DMRG techniques; namely, conventional DMRG\n(hereafter referred to simply as DMRG), dynamical DMRG\n(DDMRG), and matrix-product-state-based infinite DMRG\n(iDMRG) methods. They are used in a complementary fash-\nion to further confirm our numerical results.\nThe DMRG method is a very powerful numerical method\nfor various (quasi-)1D quantum systems57. However, some\ndifficulties are often involved in the DMRG analysis for\nstrongly frustrated systems like Eq. (1). First, the system\nsize dependence of physical quantities is usually not straight-\nforward. Therefore, relatively many data points are required\nto perform a reasonable finite-size scaling analysis. We thus\nstudy systems with length up to L= 161 (LB= 81;LA=\n80) under open boundary conditions (OBC) and systems with\nlength up to L= 64 (LB= 32;LA= 32 ) under periodic\nboundary conditions (PBC). Either OBC or PBC is chosen\ndepending on the calculated quantity. Second, a lot of nearly-\ndegenerate states are present around the ground state. To\nobtain results accurate enough, a relatively large number of\ndensity-matrix eigenstates mmust be kept in the renormal-\nization procedure. In this paper, we keep up to m= 8000density-matrix eigenstates, which is much larger than that kept\nin usual DMRG calculations for 1D systems, and extrapolate\nthe calculated quantities to the limit m!1 if necessary. In\nthis way, we can obtain quite accurate ground states within the\nerror of \u0001E=L = 10\u00008jJ1j.\nFor the calculation of dynamical quantities, we use the\nDDMRG method which has been developed for calculating\ndynamical correlation functions at zero temperature in quan-\ntum lattice models58. Since the DDMRG algorithm performs\nbest for OBC, we study a open cluster with length up to\nL= 129 (LB= 65;LA= 64 ). The DDMRG approach is\nbased on a variational principle so that we have to prepare\na ‘good trial function’ of the ground state with the density-\nmatrix eigenstates. Therefore, we keep m= 1200 to ob-\ntain the ground state in the first ten DMRG sweeps and keep\nm= 600 to calculate the excitation spectrum. In this way, the\nmaximum truncation error, i.e., the discarded weight, is about\n1\u000210\u00005, while the maximum error in the ground-state and\nlow-lying excited states energies is about 10\u00004jJ1j.\nThe iDMRG method is very useful because it enables us to\nobtain the physical quantities directly in the thermodynamic\nlimit59,60, if the matrix product state is not too complicated\nand the simulation can be performed accurately enough. In\nour iDMRG calculations, typical truncation errors are 10\u00008\nusing bond dimensions \u001fup to 6000 . In this way, the effective\ncorrelation length near criticality is less or at most equal to\n500, so that most of interesting parameter region of the system\n(1) can be reasonably examined by our iDMRG simulations.\nIII. PREVIOUS STUDIES FOR LIMITING CASES\nSo far, our system for two limiting cases, namely, J1-J2\nchain (\r= 1) and delta chain ( \r= 0), has been extensively\nstudied. In this section, we briefly summarize to-date knowl-\nedge on the ground state for the limiting cases.\nA.J1-J2chain (\r= 1)\nAt\r= 1, we are dealing with the J1-J2chain, which may\nbe also recognized as symmetricJ1-J2zigzag ladder. In the\nlimit ofJ2=jJ1j= 0, the system is a simple FM Heisenberg\nchain with FM ordered ground state. Increasing J2=jJ1j, the\nFM state persists up to J2=jJ1j= 1=4; then, a first-order\nphase transition from the FM to an incommensurate (“spiral”)\nstate occurs61. The total spin in the incommensurate phase\nis zero (Stot= 0). Since quantum fluctuations disappear at\nthe FM critical point, the critical value J2=jJ1j= 1=4can\nbe recognized similarly both in the quantum as well as in the\nclassical model62,63.\nAtJ2=jJ1j>1=4, the incommensurate correlations are\nshort ranged in the quantum model64,65. Instead, the sys-\ntem exhibits a spontaneous nearest-neighbor FM dimeriza-\ntion with breaking of translation symmetry, as a consequence\nof the quantum fluctuations typical of magnetic frustration,\ni.e, order by disorder. By regarding the ferromagnetically4\ndimerized spin-1/2 pair as a spin- 1site, the system is effec-\ntively mapped onto a spin- 1Heisenberg chain and an Affleck-\nKennedy-Lieb-Tasaki (AKLT)-like hidden topological order\nprotected by global Z2\u0002Z2symmetry is naively expected as\na Haldane state11,66(also see Sec. IV C 6). In fact, the hidden\norder has been numerically confirmed19,20.\nFurthermore, the existence of (exponentially small) singlet-\ntriplet gap at J2=jJ1j>\u00183:3was predicted by the field-theory\nanalysis67and its verification had been a longstanding issue.\nOnly recently, the gap was numerically estimated: with in-\ncreasingJ2=jJ1jit starts to open at J2=jJ1j= 1=4, reaches its\nmaximum\u00180:007jJ1jaroundJ2=jJ1j= 0:65, and exponen-\ntially decreases20. The ground state is a kind of VBS state with\nspin-singlet formations between third-neighbor sites. There-\nfore, the magnitude of gap basically scales to the strength of\nthird-neighbor valence bond.\nB. Delta chain ( \r= 0)\nAt\r= 0, the system is the delta chain consisting of a lin-\near chain of corner-sharing triangles [Fig. 1(b)]. The ground\nstate properties are less understood than those of the J1-J2\nchain. One main reason is that numerical investigation of the\ndelta chain is particularly difficult due to the strong magnetic\nfrustration and a number of nearly-degenerate states near the\nground state. Especially at the FM critical point is macroscop-\nically degenerate and consists of multi-magnon configurations\nformed by independent localized magnons and the special lo-\ncalized multi-magnon complexes48.\nNonetheless, a numerical study could identify the ground\nstate atJ2=jJ1j<1=2to be FM; that at J2=jJ1j>1=2\nto be ferrimagnetic45. The total spins of the ferromagnetic\nand ferrimagnetic phases are L=2andL=4, respectively. To\nunderstand the origin and the properties of this ferrimagnetic\nstate, the delta chain in the large limit of easy-axis exchange\nanisotropy was studied68. In this limit the system can be re-\nduced to a 1D XXZ basal chain under a static magnetic field\ndepending on the magnetic structure of apical chain. The\nground state was identified as ferrimagnetic with fully po-\nlarized apical spins and weakly polarized basal spins. It is\nexpected that some essential features may be inherent in the\nisotropic SU(2) limit. In fact, the spin structure agrees quali-\ntatively to the ferrimagnetic state determined in this paper [see\nFig. 1(c)].\nIV . RESULTS\nA. classical limit\nAs mentioned above, the FM critical point is known to be\nJ2=jJ1j= 1=4for theJ1-J2chain (\r= 1) andJ2=jJ1j= 1=2\nfor the delta chain ( \r= 0). To be examined first is how\nthe critical point changes between the limiting cases, i.e.,\n0< \r < 1. Since the quantum fluctuations vanish at the FM\ncritical point, the critical value can be exactly estimated by\nthe classical spin wave theory (SWT). The Fourier transform\n0 0.2 0.4 0.6 0.8 100.20.40.60.81\n(a)\n(b)\n0 0.2 0.4 0.6 0.8 100.20.40.60.81\n0.1\n0.4 0.6 0.800.2\n0.6 0.8FIG. 2. (a) Classical ground-state phase diagram of the asymmetric\nJ1-J2zigzag ladder. The phases are characterized by propagation\nvectorq=qminminimizingJq[Eq. (3)]: FM ( qmin= 0); commen-\nsurate (qmin=\u0019); incommensurate ( 0< qmin< \u0019). (b) Ground-\nstate phase diagram of the asymmetric J1-J2zigzag chain [Eq. (1)]\ndetermined by DMRG calculations. Inset: enlarged view around the\nPF phase.\nof our Hamiltonian (1) reads\nH=1\n2X\nqJq~Sq\u0001~S\u0000q (2)\nwith\nJq=(1 +\r)J2cosq\n\u0006q\n(1\u0000\r)2J2\n2(1\u0000cosq)2+ 4J2\n1cos2(q=2):(3)\nIf Eq. (3) has a minimum at q= 0, the system is in an FM\nground state. The FM critical point is thus derived as\nJ2;c\njJ1j=1\n2(1 +\r): (4)5\nAs shown in Fig. 2(a), the FM region is simply shrunk with\nincreasing\rbecause the AFM interaction is increased in the\napical chain. We have confirmed this FM critical boundary\nnumerically by calculating the total spin Stotof the whole\nsystem, which is defined as\nh~S2i=Stot(Stot+ 1) =X\ni;jh~Si\u0001~Sji: (5)\nIt can be also verified by finding the absence of LRO FM state\nin the spin-spin correlation functions. These results are shown\nin Appendix A.\nBy evaluating q(\u0011qmin) value to minimize Eq. (3), a\nclassical ground-state phase diagram is obtained as Fig. 2(a).\nThere are three kinds of LRO phases: FM phase with qmin=\n0, incommensurate phase 0< q min< \u0019 , and commensu-\nrate phase with qmin=\u0019. Since the ferrimagnetic state in\nFig. 1(c) is of commensurate with q=\u0019and the propagation\nnumber of the J1-J2chain is incommensurate, the SWT re-\nsults are consistent with those of the quantum system (1) in\nthe two limiting cases \r= 0 and\r= 1. Therefore, even in\nthe quantum system an incommensurate-commensurate phase\ntransition is naively expected at finite \rwithJ2=jJ1jfixed.\nB.\r= 0: delta chain\nAlthough the ground state of the delta chain is most prob-\nably ferrimagnetic at J2=jJ1j>1=2, the detailed magnetic\nstructure and properties have not been fully settled. To gain\nfurther insight into them, we here calculate the total spin, spin-\nspin correlation functions, and stabilization energy of ferri-\nmagnetic state for the delta chain. We need to pick through\nthe system-size dependence of the quantities to deal with non-\ntrivial finite-size effects of the delta chain.\n1. total spin\nInvestigating the total spin Stotis a simple way to explore\nthe possibility of ferrimagnetic state. As shown below, the\nsystem-size dependence of Stotis significantly different be-\ntween applying OBC and PBC. This obviously implies the\ndifficulty of performing numerical calculations for this sys-\ntem. Nevertheless, the fact that they should coincide in the\nthermodynamic limit L!1 makes the finite-scaling analy-\nsis even more reliable. We here use the DMRG method.\nLet us first see the case of OBC. In Fig. 3(a) the total spin\nper siteStot=Lis plotted as a function of inverse system size\n1=Lfor several values of J2=jJ1j. WhenJ2=jJ1jis order\nof1, the effect of strong frustration is explicitly embedded\nin the finite-size scaling; namely, due to the Friedel oscilla-\ntion,Stot=Lawkwardly oscillates with 1=L. However, as ex-\npected, such an oscillation no longer appears in a case of large\nJ2=jJ1j= 100 . WhenJ2=jJ1j\u001d1, a straightforward scaling\nis allowed since the frustration is much weaker. Eventually,\nfor all theJ2=jJ1jvalues,Stot=Lseems to be extrapolated to\n1=4in the thermodynamic limit. Although one may think that\n(a)\n(b)0 0.01 0.02 0.03 0.04 0.050.10.20.3\n0 0.02 0.04 0.06 0.0800.10.20.30.4FIG. 3. Finite-size scaling analysis of the total spin per site for the\ndelta chain ( \r= 0), where (a) OBC and (b) PBC are applied. The\ndotted lines are guide for eyes.\nthe PBC should be applied if the Friedel oscillation is prob-\nlematic, the situation is not so simple as explained below.\nWe then turn to the case of PBC. In Fig. 3(b) Stot=Lis plot-\nted as a function of 1=Lfor several values of J2=jJ1j. Unlike\nin the case of OBC, the oscillation is not seen in Stot=Lvs.\n1=L. Instead, there exists a ‘critical’ system size to achieve\nfiniteStot=Lin the ground state. This is caused by a kind\nof typical finite-size effects: Under the PBC, the basal chain\nforms a plaquette singlet and the basal spins are more or less\nscreened. This screening leads to the reduction of FM in-\nteraction between the basal and apical chains although the\nFM fluctuations between the two chains are essential to sta-\nbilize the ferrimagnetic state (see Sec. IV B 2). Besides, only\na small screening may be sufficient to collapse the ferrimag-\nnetic state because the stabilization gap of ferrimagnetic state\nis very small (see Sec. 5). Consequently, the ferrimagnetic\nstate can be readily prevented under the PBC. For reference,\nthe system-size dependence of energies for spin-singlet and\nferrimagnetic states is shown in Appendix B. Since the triplet\nexcitation gap of plaquette singlet roughly scales to J2=jJ1j\nwith a fixed system size, the critical system size is larger for\nlargerJ2=jJ1jas seen in Fig. 3(b). Once the system size goes\nbeyond the critical one, Stot=Lapproaches smoothly to 1=4\natL!1 . The fitting function is Stot=L= 1=4 + 1=Lfor\nanyJ2=jJ1j.6\n0 50 100\n|i−j|−0.10.00.10.2/angbracketleftBig\nSz\nα,iSz\nβ,j/angbracketrightBig\n(a)\nA-A\nA-BB-B\n0 50 100\n|i−j|−0.10.00.10.2/angbracketleftBig\nSz\nα,iSz\nβ,j/angbracketrightBig\n(b)\n0 50 100\n|i−j|−0.10.00.10.2/angbracketleftBig\nSz\nα,iSz\nβ,j/angbracketrightBig\n(c)\n0.6 1.0 1.5 2.0\nJ2/|J1|0.00.10.20.30.40.5\n/angbracketleftBig\nSz\nA,i/angbracketrightBig\n/angbracketleftBig\nSz\nB,i/angbracketrightBig(d)\nFIG. 4. Spin-spin correlation function hSz\n\u000b;iSz\n\f;jifor the delta chain\n(\r= 0) as a function of distance ji\u0000jjat (a)J2= 0:6, (b)1, and\n(c)2. The totalSzsector is set to be Sz\ntot=L=4. The legends\ndenote as A-A: (\u000b;\f) = (A;A), B-B: (\u000b;\f) = (B;B), and A-B:\n(\u000b;\f) = (A;B). (d) Averaged values of hSz\n\u000b;iion the apical and\nbasal sites as a function of J2=jJ1j.\nThus, we have confirmed that the total spin of the delta\nchain atJ2=jJ1j>1=2isStot=L=4, indicating the ferri-\nmagnetic state, and the ferrimagnetic phase persists up to the\nlargeJ2=jJ1jlimit.\n2. spin-spin correlation\nThen, in order to determine the magnetic structure of the\nferrimagnetic state, we examine spin-spin correlation func-\ntionshS\u000b;i\u0001S\f;jibetween apical sites: (\u000b;\f) = (A;A),\nbetween basal sites: (\u000b;\f) = (B;B), and between apical and\nbasal sites: (\u000b;\f) = (A;B)=(B;A). For convenience, we\nfix the z-component of total spin at the total spin retained in\nthe ferrimagnetic ground state, i.e., Sz\ntot=Stot=L=4. It\nlifts the ground-state degeneracy due to the SU(2) symmetry\nbreaking and the (spontaneous) magnetization direction is re-\nstricted to the z-axis direction. Hence, we need only to see\nhSz\n\u000b;iSz\n\f;jiinstead ofhS\u000b;i\u0001S\f;ji.\nIn Fig. 4(a)-(c) iDMRG results of the correlation function\nhSz\n\u000b;iSz\n\f;jiare plotted as a function of distance ji\u0000jjfor\nJ2=jJ1j= 0:6,1, and 2. We clearly see long-ranged cor-\nrelations indicating a magnetic order for all the J2=jJ1jval-\nues. If no magnetic order exists, all the correlation func-\ntions converge to (Sz\ntot=L)2= 1=16at long distance limit\nji\u0000jj ! 1 . The intrachain correlations hSz\nA;iSz\nA;jiand\nhSz\nB;iSz\nB;jiare both positive at the long distance, and the for-\nmer correlation is much larger than the latter one. This means\nthat the apical spins are nearly-fully polarized and the basalspins are only weakly polarized. In addition, a N ´eel-like stag-\ngered oscillation is clearly seen leastwise at the short distance\ninhSz\nB;iSz\nB;ji. These correlations immediately correspond\nto a ferrimagnetic state, which is schematically sketched in\nFig. 1(c). This picture seems to be valid in the whole region\nofJ2=jJ1j>0:5.\nNow, another question may arise: Is the N ´eel-like staggered\nspin alignment on the basal chain LRO? The answer is NO.\nAlthoughhSz\nB;iSz\nB;jiindeed exhibits an AFM oscillation at\nshort distancesji\u0000jj, it vanishes at the long distance limit.\nIn fact, the oscillating part of hSz\nB;iSz\nB;jiexhibits a power-\nlaw decay indicating a critical behavior of the Tomonaga-\nLuttinger liquid (TLL) (see Appendix C). This also excludes\nthe possibility of VBS formation in the basal chain because\nan exponential decay of the spin-spin correlation should be\nfound in a VBS state. It means that, very surprisingly, the\npresent ferrimagnetic order is an order by disorder but without\nany geometrical symmetry breaking. Instead, the magnetic\nfrustration is relaxed by a spontaneous breaking of the global\nspin-rotation symmetry. In other words, the system can gain\nenergy from the FM interaction between the apical and basal\nchains by polarizing both of the chains. This is a very rare\ntype of order by disorder.\nThis picture of order by disorder may be further convinced\nby looking at the relation between the polarization level of\nbasal spins and the stabilization of ferrimagnetic order. Thus,\nto see theJ2=jJ1j-dependence of polarization level in more\ndetail, we examine the expectation values of local spin mo-\nmentSz\nA;iandSz\nB;i. In Fig. 4(d) the averaged values of\nhSz\nA;iiandhSz\nB;iiare plotted as a function of J2=jJ1j. We\nfind that with increasing J2=jJ1j,hSz\nA;iiincreases and sat-\nurates at 1=2; while,hSz\nB;iidecreases and goes down to 0\nin the large J2=jJ1jlimit. The decrease of hSz\nB;iimay be\nnaively expected by the fact that the basal chain approaches\na 1D SU(2) Heisenberg model at large J2=jJ1j; while, the\nincrease ofhSz\nB;iiis a simple consequence of the condition\nhSz\nA;ii+hSz\nB;ii= 1=2. If the order-by-disorder picture asso-\nciated by interchain FM interaction is true, the ferrimagnetic\nstate at larger J2=jJ1jshould be fragile because of smaller po-\nlarization of basal spins. Actually, the stabilization energy of\nferrimagnetic order decreases with increasing J2=jJ1jas de-\nscribed in the following subsection.\n3. stabilization gap\nTo figure out the stability of the ferrimagnetic state, we\ncalculate a stabilization gap defined by energy difference be-\ntween the ferrimagnetic state and a lowest state in the Stot= 0\nsector:\n\u0001(L) =E0(0;L)\u0000E0(Sg:s:;L);\u0001 = lim\nL!1\u0001(L);(6)\nwhereE0(S;L)is the lowest-state energy of L-site periodic\nsystem in the Stot=Ssector, andSg:s:is a total spin of the\nground state. The energy of ferrimagnetic state can be sim-\nply obtained as a ground-state energy, though, as mentioned7\n0.5 1 1.5 21 210-310-20 0.05 0.1 0.1500.010.020.030.04\n00.0050.010.015(a)\n(b)\nFIG. 5. Energy difference between the lowest Stot= 0 state and\nferrimagnetic ground state for the delta chain, as a stabilization gap\nof the ferrimagnetic state. (a) Finite-size scaling and (b) the extrapo-\nlated values \u0001=jJ1jas a function of J2=jJ1j. Inset: Semi-log plot of\n\u0001=jJ1jas a function of J2=jJ1j. The dotted line is a fit for the large\nJ2=jJ1jregion: \u0001=jJ1j= 0:037 exp(\u00001:3J2=jJ1j).\nabove, the total spin Sg:s:=Lin the ferrimagnetic state can de-\nviate from 1=4for a finite cluster. Whereas, it is nontrivial\nto estimate the energy of a lowest state in the Stot= 0 sec-\ntor because it is an excited state and there are many quasi-\ndegenerate states near the ferrimagnetic ground state. A\nproper way needs to be provided to extract the Stot= 0state.\nWe therefore consider the following augmented Hamiltonian\nH0=H+\u0015~S2; (7)\nwhereHis our original Hamiltonian (1) and the additional\nterm corresponds to the total spin operator ( \u0015>0). By setting\n\u0015to be large enough, all the states with Stot>0are lifted.\nEventually, we can find a state with Stot= 0 as the lowest\nstate.\nIn Fig. 5(a) the energy difference \u0001(L)is plotted as a func-\ntion of 1=Lfor severalJ2=jJ1jvalues. Since the magnetic\nfrustration causes a sine-like oscillation in \u0001(L)vs.1=L20,\nthe finite-size scaling analysis is not very straightforward.\nStill, the plotted values show the sine-like oscillation with\nroughly more than one period, an acceptable scaling with lin-\near function \u0001(L)=jJ1j= \u0001=jJ1j+A=L(Ais fitting parame-\nter) may be possible. Actually, even if we assume a more gen-\neral fitting function \u0001(L)=jJ1j= \u0001=jJ1j+A=L\u0011, the extrap-\n012\n0\n00.10.20.3\n0\n00.10.20.30.4\n0\n00.10.20.3\n0(a) (b)\n(c) (d)0 1 Intensity (arb. units)FIG. 6. Dynamical spin structure factors for (a) apical and (b) basal\nchains atJ2=jJ1j= 0:6. (c)(d) The same spectra at J2=jJ1j=\n1. Finite broadening \u0011is introduced: \u0011= 0:03jJ1jin (a) and (c),\n\u0011= 0:02jJ1jin (b), and\u0011= 0:05jJ1jin (d). The dotted lines\nare approximate analytical expressions of the main dispersions (see\ntext).\nolated values of \u0001=jJ1jare almost unchanged because \u0011\u00191\nis always achieved. In Fig. 5(b) the extrapolated values of \u0001\nare plotted as a function of J2=jJ1j. At the FM critical point\nJ2=jJ1j= 1=2, the lowest energies for all Stotsectors are de-\ngenerate and it leads to \u0001 = 0 . As soon as the system goes\ninto the ferrimagnetic phase, the stabilization gap \u0001steeply\nincreases, reaches a maximum around J2=jJ1j= 1, and de-\ncreases with further increasing J2=jJ1j. The magnetic frustra-\ntion would be strongest at the maximum position J2=jJ1j\u00181\nbecause each triangle is fully frustrated. This is another in-\ndication of the fact that the ferrimagnetic state is originated\nfrom order by disorder. As shown in the inset of Fig. 5(b), the\nstabilization gap seems to decay exponentially with J2=jJ1jin\nthe largeJ2=jJ1jregion. It means that the ferrimagnetic state\nis rapidly destabilized in the large J2=jJ1jregime although it\npersists up to J2=jJ1j=1in a precise sense. This is con-\nsistent with the rapid decrease of hSz\nB;iiwithJ2=jJ1j. The\nsystem can gain only little energy from the interchain FM in-\nteraction in case where the basal spins are not really polarized.\nIn short, the quantum fluctuations between the apical and the\nbasal chains play an essential role to stabilize the ferrimag-\nnetic state.\n4. dynamical spin structure factor\nIn order to provide further insight into the ferrimagnetic\nstructure, we investigate the low-energy excitations of the8\ndelta chain. We calculate dynamical spin structure factors\nfor both the apical and basal chains with using the DDMRG\nmethod. The dynamical spin structure factor is defined as\nS\u000b(q;!) =1\n\u0019Imh 0jSz\n\u000b;q1\n^H+!\u0000E0\u0000i\u0011Sz\n\u000b;qj 0i\n=X\n\u0017jh \u0017jSz\n\u000b;qj 0ij2\u000e(!\u0000E\u0017+E0); (8)\nwhere\u000bdenotes either apical (A) or basal (B) chain, j \u0017iand\nE\u0017are the\u0017-th eingenstate and the eigenenergy of the system,\nrespectively ( \u0017= 0 corresponds to the ground state). Under\nOBC, we define the momentum-dependent spin operators as\nSz\n\u000b;q=r\n2\nL\u000b+ 1X\nleiqriSz\n\u000b;i; (9)\nwith (quasi-)momentum q=\u0019Zx=(L\u000b+ 1) for integers 1\u0014\nZx\u0014L. We use open clusters with LA= 31;LB= 32\nforSB(q;!)and withLA= 64;LB= 65 forSA(q;!). In\nFig. 6, DDMRG results of the dynamical spin structure factors\nare shown for J2=jJ1j= 0:6and1.\nLet us see the spectrum for the apical chain. Since the\nspins are fully polarized on the apical chain, the main dis-\npersion ofSA(q;!)is basically described by that of the 1D\nFM Heisenberg chain. For J2=jJ1j= 0:6, a precise fitting\nleads to!q=jJ1j=J0(cos(q)\u00001) +J00(cos(2q)\u00001)with\nnearest-neighbor FM coupling J0=\u00000:11and next-nearest-\nneighbor AFM coupling J00= 0:016. Interestingly, we find\nthat the dominant FM coupling is effectively induced on the\napical chain in spite of no direct interaction between apical\nsites. A similar fitting for J2=jJ1j= 1 givesJ0=\u00000:075\nJ00= 0:018. The reduction of jJ0jwith increasing J2=jJ1j\nis naturally expected because the apical spins are completely\nfree in the large J2=jJ1jlimit. This reduction also reflects the\nweakening of ferrimagnetic state.\nWe then turn to the spectrum for the basal chain. Although\nthe basal chain is weakly polarized, we expect that the funda-\nmental excitations could be at least qualitatively described by\nthose of the 1D SU(2) Heisenberg chain because the dom-\ninant short-range correlation is AFM. For J2=jJ1j= 0:6,\nthe magnon dispersion (lower bound of the continuum) of\nSB(q;!)is certainly sine-like function and the well-known\nshaped two-spinon continuum is seen. However, surpris-\ningly, such the weak spin polarization ( SB;tot=LB= 0:11)\ndrastically suppresses the dispersion width down to 0:15jJ1j\nfrom that of the 1D SU(2) Heisenberg chain \u0019jJ1j=269. It\nis interesting that the dispersion width is rapidly recovered\nto0:75jJ1jwhen the spin polarization is slightly reduced to\nSB;tot=LB= 0:085atJ2=jJ1j= 1. Another effect of the\nweak polarization on the dispersion is a shift of node. Due to\nthe dominant AFM fluctuation on the basal chain in the whole\nregion of ferrimagnetic phase, a largest peak always appears\nat(q;!) = (\u0019;0). If there is no spin polarization, the other\nnode should be at q= 0 but it is actually shifted to higher q\nvalue as seen in Fig. 6(b)(d). This behavior is similar to the\ncase in the presence of magnetic field. Namely, the node po-\nsition can be expressed as q= 2\u0019hSB;toti=LBwherehSB;toti\nis the total spin of the basal chain with length LB.\n0 0.01 0.02 0.0300.10.20.3\n0 0.1 0.200.10.20.3(a)\n(b)FIG. 7. (a) System-size dependence of total spin per site Stot=Lfor\nseveral\rvalues withJ2=jJ1j= 0:6fixed. The dashed line indicates\nthe value ofStot=Lfor the full ferrimagnetic state. (b) The L!1\nextrapolated values of Stot=Las a function of \r.\nC. Finite\r: asymmetric J1-J2zigzag ladder\nAs described above, we have confirmed that the ferrimag-\nnetic state is indeed stabilized at J2=jJ1j>1=2in the delta\nchain (\r= 0). Then, let us see what happens when the apical\nsites are connected by AFM interaction, the strength of which\ncan be controlled by \r. Since the system is in a singlet ground\nstate, i.e.,Stot= 0, in theJ1-J2chain (\r= 1), the collapse of\nferrimagnetic state is naively expected at some \r(<1). Inci-\ndentally, the ferrimagnetic state is trivially enhanced if an FM\ninteraction is introduced between apical sites.\n1. total spin\nSimply, we examine the \r-dependence of the total spin\nto identify when and how the ferrimagnetic state is destabi-\nlized. In Fig. 7(a) the total spin per site Stot=Lis plotted as\na function of 1=Lfor several\rvalues with J2=jJ1j= 0:6\nfixed. Due to the strong frustration, the value of Stot=Los-\ncillates with 1=L; however, we may perform a reasonable\nfinite-size scaling analysis with finer data points using large\nenough clusters. We here use open clusters with length up to\nL=LA+LB= 100+101 = 201 . TheL!1 extrapolated9\nvalue ofStot=Lis plotted as a function of \rin Fig. 7(b).\nAt\r= 0, the ground state is in the ferrimagnetic state\nwith nearly-fully polarized apical spins and the total spin per\nsite isStot=L=4. Hereafter, we call this state “ fullferri-\nmagnetic (FF) state” to discriminate it from another ferrimag-\nnetic state with Stot< L= 4(denoted as “ partial ferrimag-\nnetic (PF) state”) which appears below. When the AFM inter-\naction between apical spins is switched on, one may naively\nexpect a collapse or weakening of the full spin polarization in\nthe apical chain; nevertheless, interestingly, the FF condition\nStot=L=4survives up to \r\u00190:08. This can be interpreted\nas follows: Roughly speaking, since the system can gain more\nenergy from interchain FM interaction than AFM interaction\nbetween apical spins, the FM LRO in the apical chain is still\nmaintained at \r<\u00180:08.\nWith increasing \rfrom 0:08, the competition between in-\nterchain FM interaction and apical intrachain AFM interaction\nderives a new state. As seen in Fig. 7(a), at \r= 0:09the value\nofStot=Lseems to converge at a finite but smaller value than\n1=4as1=Ldecreases. This clearly suggests that some sort\nof collapse of the FF state happens around \r= 0:09. With\nfurther increasing \r, the value of Stot=Lappears to be con-\ntinuously reduced and reaches zero around \r= 0:14[see\nFig. 7(b)]. Surprisingly, we find that there exists a finite \r-\nrange exhibiting 0< S tot=L < 1=4. Since the basal chain\nbasically keeps its weakly polarized or nearly singlet state,\nit would be a good guess that spin polarization on the api-\ncal sites is gradually collapsed with increasing \rin this PF\nphase ( 0:08<\u0018\r<\u00180:14). In other words, the ferrimagnetic\norder by disorder in association with the global spin-rotation-\nsymmetry breaking disappears around \r= 0:14. Actually, as\nstated below, the system has the other order by disorder, i.e.,\ndimerization order, at \r>\u00180:14. The region of the PF phase\nis shown as a shaded area in the ground-state phase diagram\n[Fig. 2(b)].\nWe make some remarks on the existence of PF phase. Such\na ‘halfway’ magnetization 0< S tot=L < 1=4in a ferrimag-\nnetic state is generally prohibited by the Marshall-Lieb-Mattis\n(MLM) theorem70,71. There is an exception to this, however,\nwhen the ferrimagnetic order and a quasi-long-range order of\nTLL compete72. This corresponds to the competition between\nsmall FM polarization and dominant AFM fluctuations in the\nbasal chain of our system. As confirmed in Appendix C, the\nbasal chain in the FF state indeed exhibits a TLL behavior.\n2. spin-spin correlation\nWe then consider the evolution of spin-spin correlation\nfunction with \rin the FF phase. In Fig. 8(a)-(c) iDMRG\nresults of the spin-spin correlation function hSz\n\u000b;iSz\n\f;jifor\n\r= 0,0:04, and 0:08with fixedJ2=jJ1j= 0:6are plotted\nas a function of distance ji\u0000jj. We keepSz\ntot= 4=Las done\nin Fig. 4.\nAs far as the FF state is maintained up to \r\u00190:08, the cor-\nrelation functions seem to be almost independent of \r. Ac-\ncordingly, the expectation values of Sz\nA;iandSz\nB;iare un-\nchanged up to \r\u00190:08, as shown in Fig. 8(d). Perhaps, one\nFIG. 8. Spin-spin correlation function hSz\niSz\njiof the asymmetric\nJ1-J2zigzag ladder as a function of ji\u0000jjwith fixedSz\ntot=L= 1=4\nandJ2=jJ1j= 0:6for (a)\r= 0, (b)0:04, and (c) 0:08. (d) Averaged\nvalues ofhSz\niion the apical and basal sites as a function of \rfor\nJ2=jJ1j= 0:6. The circles and crosses denote iDMRG and DMRG\nresults, respectively.\nmay naively expect the reduction of hSz\nA;iiwith increasing\n\r, i.e., with increasing AFM coupling between apical spins.\nHowever, this is not actually the case. This can be understood\nas follows: as estimated by the fitting of low-energy excitation\nspectrum, an FM coupling with the magnitude \u00180:11jJ1jis\neffectively induced between neighboring apical sites at \r= 0.\nThus, the apical chain may be effectively mapped onto an FM\nchain withJe\u000b=\u00000:11jJ1j. Therefore, it would be rather\nnatural that the (nearly) full polarization is free of the influ-\nence of additional AFM coupling \rJ2until it reaches around\n\u00180:11jJ1j.\nNo\r-dependence of the spin structure up to \r\u00190:08then\nindicates that the total energy of FF state is simply lifted by\nthe AFM interaction \rJ2between nearly-fully polarized api-\ncal spins; it switches into the energy level with a metastable\nPF state at\r\u00190:09. Hence, the FF to PF phase transition\nis of the first order. It can be also confirmed by a steep (or\nalmost discontinuous) change of StotandhSz\n\u000b;iiat\r\u00190:09.\nOn the other hand, both of hSz\nA;iiandhSz\nB;iismoothly ap-\nproach to zero around \r\u00190:14. Thus, the transition from PF\nto the spiral singlet ( Stot= 0) region is of the second order or\ncontinuous.\n3. stabilization gap\nTo quantify the stability of ferrimagnetic state at finite \r,\nwe calculate the stabilization energy \u0001[Eq.(6)] as done in\nthe case of delta chain. In Fig. 9(a) the finite-size scal-10\n0 0.05 0.1 0.1500.010.020.03\n0 0.02 0.04 0.06 0.08 0.100.0020.0040.0060.008(a)\n(b)\nFIG. 9. (a) Finite-size scaling of energy difference between lowest\nStot= 0 state and ferrimagnetic ground state for the asymmetric\nJ1-J2zigzag ladder with fixed J2=jJ1j= 0:6. The solid and dotted\nlines show the fitting results with \u0001(L)=jJ1j= \u0001=jJ1j+ A=Land\n\u0001(L)=jJ1j= \u0001=jJ1j+ A=L\u0011, respectively. (b) Extrapolated val-\nues of \u0001=jJ1jas a function of \r. The width of error bar means the\ndifference of \u0001=jJ1jobtained by the two fitting functions.\ning analysis of \u0001(L)is performed for several \rvalues with\nJ2=jJ1j= 0:6fixed. Although a sine-like oscillation is\npresent as in the case of delta chain, an acceptable scaling may\nbe possible in the FF phase ( \r<\u00180:09). We here employ two\nkinds of fitting functions: \u0001(L)=jJ1j= \u0001=jJ1j+ A=Land\n\u0001(L)=jJ1j= \u0001=jJ1j+ A=L\u0011. Since the former (latter) func-\ntion seems to underestimate (overestimate) the extrapolated\nvalue of \u0001=jJ1j, their averaged value is plotted as a function\nof\rin Fig. 9(b). The width of error bar means the differ-\nnce between two values of \u0001=jJ1jobtained by the two fitting\nfunctions. The stabilization gap is approximately linearly re-\nduced by\rup to the critical point \r\u00190:08. This also sup-\nports the above speculation that the total energy of FF state is\nsimply lifted by \rJ2.\nIn the PF phase ( \r>\u00180:09), however, the scaling analysis\nof\u0001(L)is virtually impossible. As an example, \u0001(L)=jJ1j\nvs.1=Lfor\r= 0:1is shown in Fig. 9(a). This difficulty is\ncaused by the following several factors: (i) As shown in the\nnext subsection, an incommensurate oscillation is involved in\nthe PF state. (ii) The total spin per site Stot=Lis strongly de-\npendent on system size since the states in different Stotsectors\nare extremely quasi-degenerate around the ground state. (iii)\n0 0.1 0.2 0.30basal cha/g3444 n\nap/g3444 cal cha/g3444 n\n0 0.2 0.4 0.6 0.8 10\n0.06 0.1 0.14-12.7-12.6(a)\n(b)FIG. 10. (a) Static spin structure factor for J2=jJ1j= 0:6as a\nfunction of\r. The lattice spacing ais set as shown in Fig. 1(a). (b)\nEnlarged figure of (a) for 0\u0014\r\u00140:3. Inset: Ground-state energy\nas a function of \r.\nThe available system size is strictly limited because the sec-\nond term of Eq.(7) includes long-range interactions and a pe-\nriodic cluster must be used. Nevertheless, the stabilization gap\nshould be positive due to the nonzero total spin of the ground\nstate [Fig. 7(b)]. We can, at least, confirm that the PF order is\nvery fragile with the stabilization gap \u0001<4:3\u000210\u00004jJ1jat\n\r= 0:08. This small stabilization gap also tells us that there\nare a macroscopic number of quasi-degenerate states belong-\ning to different Stotsectors, since the total spin is continu-\nously varied from L=4to0with\rin the PF phase.\n4. Static spin structure factor\nIt is important to see how the intrachain spin modulation\nchanges with \r. A most significant quantity to know it would\nbe the propagation number which can be extracted as a max-\nimum position of static spin structure factor. Therefore, we\ncalculate the static spin structure factor for each of the apical\nand basal chains. It is defined as\nS\u000b(q) =1\nL2\u000bL\u000bX\ni:j=1eiq(ri\u0000rj)hS\u000b;i\u0001S\u000b;ji (10)\nfor the apical ( \u000b= A ) or basal (\u000b= B) chains. The lattice\nspacingais set as shown in Fig. 1(a). In Fig. 10 the propaga-11\ntion numbers qmaxforJ2=jJ1j= 0:6using a 64-site periodic\ncluster are plotted as a function of \r.\nAt\r= 0, the system is in the FF state whose spin structure\nis commensurate and simple as shown in Fig. 1(c). The apical\nspins are nearly-fully polarized and the propagation number\nisqmax= 0, whereas for basal chain the dominant correlation\nis AFM (qmax=\u0019=a) although it is slightly polarized. It is\nobvious that this spin structure persists in the whole region of\nthe FF phase at 0<\u0018\r<\u00180:08.\nIn the singlet ( Stot= 0) phase at\r>\u00180:14, the propagation\nnumbers of apical and basal chains are both incommensurate,\ni.e.,0< q max< \u0019=a . At\r= 1, they coincide and it is\nestimated as qmax= 0:958. This value is in good agreement\nwith our previous estimation qmax= 0:95873in the thermo-\ndynamic limit. With decreasing \r, the propagation number is\nreduced because short-range FM correlation is relatively en-\nhanced. Interestingly, they are equal or very close down to the\ncritical point at \r\u00190:14and obviously split for smaller \r. It\nwould be a good guess that the short-ranged spiral structure of\ntheJ1-J2chain (\r= 1) is approximately maintained down to\n\r\u00190:14. In a broad sense, this incommensurate region can\nbe referred to as a spiral singlet phase.\nIn Fig. 10(b) an enlarged figure around the PF phase\n(0:08<\u0018\r<\u00180:14) is shown. As shown in the inset of\nFig. 10(b) the phase boundaries are recognized by level cross-\ning of the ground state energies. It clearly indicates the exis-\ntence of an intermediate phase between the FF and spiral sin-\nglet phases, though the region ( 0:08>\u0018\r>\u00180:11) is a bit nar-\nrower than that in the thermodynamic limit ( 0:08>\u0018\r>\u00180:14)\ndue to finite size effects. The intermediate phase is the PF\nphase as described above.\nIn the PF phase, the dominant correlation of the basal chain\nseems to be incommensurate and the propagation number of\nthe apical chain keeps qmax= 0. It is a natural consequence of\nthe global spin-rotation-symmetry breaking because the total\nspin can be no longer finite if both of the propagation num-\nbers are nonzero. This incommensurate propagation is a con-\nsequence of the halfway magnetization so that the prohibition\nby the MLM theorem is also avoided by the TLL characteris-\ntic of basal chain. Similar incommensurate features have been\nreported for PF state in frustrated systems74,75.\n5. Dimerization order\nSo, let us see more about the magnetic properties of the\nspiral singlet ( Stot= 0) phase at larger \r. It is known that\nthe system has LRO with spontaneous dimerization in the J1-\nJ2chain (\r= 1)19,20. Therefore, as a starting point it would\nbe reasonable to examine the evolution of dimerization order\nparameters with decreasing \rfrom 1. The dimerization order\nparameter between sites distant \u000ealong theJ1zigzag chain is\ndefined as\nOdimer(\u000e) = lim\nL!1jhSA;i\u0001SB;i\u0000(\u000e\u00001)=2i\n\u0000hSA;i\u0001SB;i+(\u000e+1)=2ij;(11)\nδ\nδ = \nδ = \nδ = (a)δ = 1\nδ = 2\nδ = 3\n(b)FIG. 11. (a) Schematic pictures of possible dimerization order.\nThe states for \u000e= 2 and\u000e= 3 are characterized as VBS. A solid\n(dotted) ellipse denotes a spin-singlet (spin-triplet) dimer. (b) Aver-\naged dimerization order parameters Odimer(\u000e)as a function of \rat\nJ2=jJ1j= 1. Inset: each contribution to Odimer(2)from the apical\nand basal chains.\nfor odd\u000e, and\nOdimer(\u000e) = lim\nL!1jhS\u000b;i\u0001S\u000b;i+\u000e=2i\u0000hS\u000b;i\u0001S\u000b;i+\u000e=2ij;\n(12)\nfor even\u000e. IfOdimer(\u000e)is finite for\u000e, it signifies a long-range\ndimerization order associated with mirror-symmetry breaking\nfor odd\u000eor translation-symmetry breaking for even \u000e. We\nhere study the case of \u000e= 1,2, and 3. Schematic pictures of\nthe possible dimerization orders are shown in Fig. 11(a). A\nfiniteOdimer(\u000e)for\u000e= 2and3indicates a valence bond for-\nmation, i.e., spin-singlet formation, between two sites on the\ndimerized bond. In Fig. 11(b) the iDMRG results of dimeriza-\ntion order parameter Odimer(\u000e)for\u000e= 1,2, and 3are plot-\nted as a function of \ratJ2=jJ1j= 1. For confirmation, we\nalso estimateOdimer(\u000e)in the thermodynamic limit for some\n\rvalues using DMRG under OBC. We can find their excel-\nlent agreement with the iDMRG results. Note that the FM\ninteractionJ1at both edges in the open clusters is set to be\nzero. It enables us to perform the finite-size scaling analysis\nmore easily because the competing two translation-symmetry\nbreaking states are explicitly separated20. For confirmation,\nwe have checked that the ground state in the thermodynamic\nlimit does not depend on the choice of boundary conditions.\nAt\r= 1, two dimerization orders with \u000e= 1 and\u000e= 3\ncoexist20. With decreasing \r, interestingly,Odimer(3)is sig-\nnificantly enhanced and Odimer(1)is slightly increased down12\nto\r\u00190:195. With fixed J2=jJ1j= 1, we may deduce that\nthe magnetic frustration is largest in the limit of \r= 0where\nthe system is a series of isotropic triangles with uniform mag-\nnitude of interactions. The valence bond pair may be strength-\nened to screen spins more strongly for the relaxation of larger\nmagnetic frustration at smaller \r. SinceOdimer(2) = 0 down\nto\r\u00190:195, this state is dominantly characterized as a VBS\nstate with\u000e= 3 dimerization order (we call it “ D3-VBS”\nstate).\nEven more surprisingly, Odimer(2)exhibits a steep increase\n(almost jump) at \r\u00190:195. The other order parameters\nOdimer(1)andOdimer(3)are also not differentiable with \r\nat this point. This clearly indicates another first-order tran-\nsition at\r\u00190:195. We note that both of the apical and\nbasal chains are spontaneously dimerized along the chain di-\nrection in the region of Odimer(2). The order parameter for\neach chain is plotted in the inset of Fig. 11(b) (The main fig-\nure shows the averaged value). Since the value of Odimer(2)\nis much larger than the other dimerization order parameters at\n0:035<\u0018\r<\u00180:195, the state is recognized as a VBS one with\n\u000e= 2 dimerization order (we call it “ D2-VBS” state). Since\nthis\u000e= 2dimerization order is associated with a translation-\nsymmetry breaking, the magnetic structure consists of a su-\npercell with four sites. More detailed analysis is given in Ap-\npendix D.\nThe dimerization order can be also detected by studying\nthe topological properties of the system. Then, let us see\nthe entanglement spectrum (ES)76which can be obtained by\na canonical representation of the an infinite matrix-product-\nstate in the iDMRG calculations12. Using Schmidt decompo-\nsition, the ground-state wave function can be expressed as\nj i=X\nie\u0000\u0018i=2j\u001eA\nii\nj\u001eB\nii; (13)\nwhere the statesj\u001eS\niicorrespond to an orthonormal basis\nfor the subsystem S(either A or B). We study the ES for\nseveral kinds of splitting pattern between subsystems A and\nB. The splitting patterns are sketched in Fig. 12(a). In our\niDMRG calculations, the ES f\u0018\u000bgis simply obtained as \u0018\u000b=\n\u00002 ln\u0015\u000b, wheref\u00152\n\u000bgare the singular values of the reduced\ndensity matrices after the bipartite splitting. The low-lying\nfour ES levels are plotted as function of \rin Fig. 12(b)-(e).\nWhen the one dimerized singlet pair straddles the subsys-\ntems A and B [Fig. 12(b) and (d)], the lowest entanglement\nlevel is doubly degenerate as a reflection of the edge state.\nOn the other hand, when this is not the case [Fig. 12(c) and\n(e)], the lowest entanglement level is non-degenerate. Our re-\nsult for theD3-VBS state is consistent with previous research\non the symmetric case19. These facts would strongly support\nthe formation of long-range dimerization order. We can also\nfind a discontinuous change of @(\u00002 ln\u0015\u000b)=@\r at\r\u00180:195,\nwhich seems to correspond to the transition point from the\nD2-VBS toD3-VBS state. We note that the difference be-\ntween (b) and (d) as well as (c) and (e) in Fig. 12 comes from\nthe asymmetric nature of our system. More details about the\nasymmetric nature are discussed in Appendix D.\nIn the spiral singlet phase, the system is in either D2-VBS\norD3-VBS state. The phase boundary between them is shown\n(a)\n(e)(b) (c)\n(d)(b) (c) (d) (e)D2-VBS\nD3-VBS\nγ γ21210121\nα=λ λFIG. 12. (a) Schematic pictures of considered splitting of the sys-\ntem into two subsystems in the D2-VBS andD3-VBS state. A solid\nellipse denotes a spin-singlet pair. The number of singlet pair cross-\ning with each cut is shown in the green square. (b)-(e) Entangle-\nment spectrum for the corresponding splitting as a function of \rat\nJ2=jJ1j= 1.\nin Fig. 2(b). In general, the spin gap, namely, energy differ-\nence between spin-singlet ground state and spin-triplet first\nexcited state, is expected to be finite when the system is in a\nVBS state. In theD3-VBS phase, the spin gap simply scales to\nan energy to break a valence bond for \u000e= 3. In theD2-VBS\nphase, each of the apical and basal chains has a different va-\nlence bond. Nevertheless, it is easy to imagine that the valence\nbond in the apical chain is more fragile because of smaller\nAFM interaction, although Odimer(2)for the apical chain is\nlarger than that for the basal chain. Thus, the spin gap in the\nD2-VBS phase scales to an energy to break a valence bond in\nthe apical chain. This means that a larger energy than the spin\ngap is needed to break a valence bond in the basal chain. It\nwould provide a 1/2-plateau in the magnetization process with\nmagnetic field.13\n0.2 0.4 0.6 0.8 1.0\nγ0.000.010.020.030.040.050.060.070.08Ostring\nFIG. 13. String order parameter as a function of \ratJ2=jJ1j= 1\nusing iDMRG (circles) and DMRG (crosses) methods. The DMRG\nresults are extrapolated values to the thermodynamic limit.\n6. String order\nWe have confirmed the existence of nearest-neighbor ( \u000e=\n1) FM dimerization order in the whole spiral singlet region. A\nspin-triplet pair may be effectively formed in the each ferro-\nmagnetically dimerized bond: By relating three states j\"\"i ,\nj\"#i +j#\"i )=p\n2, andj##i toSz= 1,0, and\u00001states,\nrespectively, the resultant spin on the dimerized bond can be\nreduced to a spin- 1degree of freedom. Consequently, the sys-\ntem could be mapped onto a S= 1 Heisenberg chain ac-\ncompanied by the emergent effective spin- 1degrees of free-\ndom with the dimerized two spin-1\n2’s19. Furthermore, the\npresence of third-neighbor AFM dimerization order ensures\na valence bond formation between the neighboring effective\nS= 1 sites20. It leads to finite spin gap as a Haldane gap\nin symmetry-protected VBS state11,12. Although the spin gap\nis a good indicator to measure the stability of VBS state, it\nwould be too small to correctly estimate with DMRG method\nin most of the parameter region of system (1). Alternatively,\nthe stability of VBS state associated with the Haldane picture\ncan be evaluated by examining the string order parameter77:\nOz\nstring =\u0000lim\njk\u0000jj!1h(Sz\nA;k+Sz\nB;k)\nexp[i\u0019j\u00001X\nl=k+1(Sz\nA;l+Sz\nB;l)](Sz\nA;j+Sz\nB;j)i:\n(14)\nFor our system (1), Eq. (14) can be simplified as\nOz\nstring =\u0000lim\njk\u0000jj!1(\u00004)j\u0000k\u00002h(Sz\nA;k+Sz\nB;k)\nj\u00001Y\nl=k+1Sz\nA;lSz\nB;l(Sz\nA;j+Sz\nB;j)i:i (15)\nThe finite value of Oz\nstring suggests the formation of a VBS\nstate having a hidden topological long-range string order.Since two-fold degeneracy of the ground state due to the FM\ndimerization is lifted in our numerical calculations, jOz\nstringj\ncan have two different values depending on how to select k\nandj. We then take their average.\nIn Fig. 13 iDMRG and DMRG results for the string order\nparameter are plotted as a function of \ratJ2=jJ1j= 1. We\ncan see a good agreement between the iDMRG and DMRG\nvalues. With decreasing \rfrom 1,Oz\nstring is significantly\nincreased and has a pointed top at \r\u00190:195, which is\nthe first-order transition point between D3-VBS andD2-VBS\nphases. With further decreasing \r, it decreases and vanishes\nat\r= 0:035, which is the second-order transition point be-\ntweenD2-VBS and ferrimagnetic phases. We notice that the\noverall trend ofOz\nstring is similar to that of the \u000e= 3 dimer-\nization order parameter Odimer(3). It means that the stability\nof string order is dominated by the strength of valence bond\nwith\u000e= 3. In other words, a Haldane state is produced as\na structure where all the neighboring effective spin- 1sites are\nbridged by the \u000e= 3 valence bonds. Even though it is a\nHaldane state, the maximum value Oz\nstring\u00180:065is much\nsmaller thanOz\nstring =4\n9'0:4444 for the perfect VBS state\nfor the AKLT model11andOz\nstring'0:3743 for theS= 1\nHeisenberg chain78. This small value of Oz\nstring is interpreted\nas a sign of fragility of the D3-VBS state, which is however\ncomparable with the maximum value for the J1-J2chain at\nJ2=jJ1j\u00190:6(Oz\nstring\u00180:06)19,20.\nV . SUMMARY\nWe studied the asymmetric S=1\n2J1-J2zigzag ladder,\ndefined as two different AFM Heisenberg chains coupled by\nzigzag-shaped interchain FM interaction, using the DMRG-\nbased techniques. The AFM chain with larger (smaller) inter-\naction is referred to as apical (basal) chain.\nFirst, a classical phase diagram was obtained by the spin-\nwave theory. It contains three phases: FM, commensu-\nrate, and incommensurate phases. It offers the possibility of\ncommensurate-incommensurate phase transition by tuning the\nratio of AFM interaction of the apical and basal chains in the\nquantum case.\nNext, we revisited the ferrimagnetism in the so-called delta\nchain as the vanishing limit of AFM interaction in the apical\nchain. The ferrimagnetic state is characterized by total spin\nStot=L=4. By carefully examining the long-range spin-spin\ncorrelation functions and low-energy excitations, we pointed\nout that the origin of ferrimagnetic state is order by disor-\nder without geometrical symmetry breaking but with a global\nspin-rotation-symmetry breaking. Accordingly, the system\ncan gain energy from the FM interaction between the polar-\nized apical and basal spins. So to speak, FM fluctuations play\nan essential role to lower the ground state energy against the\nmagnetic frustration. This is a rare type of order by disor-\nder. And yet the basal chain is essentially a critical AFM\nHeisenberg chain as a TLL and its polarization is rather ill-\nconditioned as a state of chain itself. In this regards, one could\ninterpret this to mean that the ferrimagnetic order competes14\nwith a quasi-long-range AFM order of TLL.\nThen, we examined how the ferrimagnetic state is affected\nby AFM interaction of the apical chain, which is controlled by\n\r. We found that the ferrimagnetic state with Stot=L=4is\nmaintained up to a finite value of \r; and with further increas-\ning\rthe system goes into spiral singlet ( Stot= 0) phase at a\ncertain amount of \r. Of particular interest is the appearance of\nanother ferrimagnetic phase characterized by 0