diff --git "a/ferromagnetic resonance/3.json" "b/ferromagnetic resonance/3.json" new file mode 100644--- /dev/null +++ "b/ferromagnetic resonance/3.json" @@ -0,0 +1 @@ +[ { "title": "2202.02461v2.Gyromagnetic_bifurcation_in_a_levitated_ferromagnetic_particle.pdf", "content": "Gyromagnetic bifurcation in a levitated ferromagnetic particle\nT. Sato,1T. Kato,1Daigo Oue,2, 3, 4and M. Matsuo2, 5, 6, 7\n1Institute for Solid State Physics, University of Tokyo, Kashiwa 277-8581, Japan\n2Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing, 100190, China.\n3The Blackett Laboratory, Imperial College London, London SW7 2AZ, United Kingdom\n4Instituto Superior T´ ecnico, University of Lisbon, 1049-001 Lisboa, Portugal\n5CAS Center for Excellence in Topological Quantum Computation,\nUniversity of Chinese Academy of Sciences, Beijing 100190, China\n6Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, 319-1195, Japan\n7RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan\n(Dated: September 20, 2023)\nWe examine mechanical rotation of a levitated magnetic particle that is induced by ferromagnetic\nresonance under microwave irradiation. We show that two stable solutions appear in a certain range\nof parameters by bifurcation when the rotation frequency is comparable to the microwave frequency.\nThis phenomenon originates from the coexistence of the Barnett and the Einstein-de Haas effects.\nWe also reveal that this measurement is sensitive to the strength of the spin-rotation coupling. Our\nwork provides a platform for accessing a microscopic relaxation process from spin to macroscopic\nrotation.\nIntroduction.— The discovery of gyromagnetism [1, 2],\ni.e., the interconversion between spin and mechanical ro-\ntation, was a milestone in magnetism research, because\nit revealed the origin of magnetism to be the intrinsic\nangular momentum of electrons, which classical physics\ncannot explain, even before the establishment of quan-\ntum mechanics. Originally, gyromagnetic effects were in-\nvestigated in bulk magnetic materials with aim of deter-\nmining their gyromagnetic ratios [3]. Recently, gyromag-\nnetic effects have been recognized as universal phenom-\nena, as demonstrated in various systems at scales ranging\nfrom those of condensed matter [4–16] to those of particle\nphysics [17–21]. They have also provided powerful tools\nwith which to measure and control both the mechanical\nand magnetic degrees of freedom [4, 5, 22–27].\nIndeed, the Barnett effect (conversion of angular mo-\nmentum of mechanical rotation into spin [1]) has been\nutilized to identify the angular momentum compensa-\ntion temperature of ferrimagnets [24, 25] and to gener-\nate spin current from rigid-body rotation [6–8], surface\nacoustic waves [13–15], and vorticity in fluids [9–12]. On\nthe other hand, the Einstein–de Haas effect (mechanical\ntorque generated from spin polarization [2]), which is ab-\nbreviated as the EdH effect hereafter, has been utilized to\nmeasure the faint torque caused by a single electron spin\nflip [23], identify the gyromagnetic ratio of a nanomag-\nnetic thin film [22], and reveal the demagnetization pro-\ncess in ferromagnets on sub-picosecond time scales [26].\nVery recently, it has been demonstrated, in a solid-\nstate device, that the Barnett and EdH effects can coexist\nin the GHz-frequency regime [16], although the Barnett\nand EdH effects have been treated independently so far.\nThrough these two routes, angular momenta can bounce\nback and forth between magnetic and mechanical degrees\nof freedom in a single system; hence, their coexistence\nmay bring out rich physics and possibly highlight themicroscopic features of the spin-rotation coupling, the\nfundamental coupling between spin and mechanical an-\ngular momentum of a rotating body [28, 29]. However,\nsolid-state platforms involve various excitations; hence,\nthe gyromagnetic phenomena are damped by and could\neven be buried in, e.g., spin and charge transport and im-\npurity scattering. In this sense, it is crucial to construct\na platform that is detached from those relaxation paths\nand enables high-frequency, stable rotation.\nLevitated optomechanics, inspired by laser tweezers,\nlevitation, and cooling [30–34], can provide such a plat-\nform. Stable levitation of small particles has been demon-\nstrated under a high vacuum through the use of, e.g.,\noptical and radio-frequency forces [35]. In particular,\nwith optical forces controlled by parametric feedback,\nsub-kelvin cooling of center-of-mass motion of small par-\nticles has been demonstrated [36–38]. By combining the\nfeedback cooling scheme with the cavity cooling tech-\nnique [39, 40], even zero-point fluctuation of the center of\nmass has been revealed [41]. Since motion of the center\nof mass can be significantly suppressed, high-frequency\nrotation ( ∼GHz) of small particles in a high vacuum\nhas been studied recently [42, 43]. The levitation of fer-\nromagnets, which is necessary for our setup, has been\nstudied theoretically in previous works [44–52] and has\nachieved also experimentally [53, 54]. Though the levi-\ntated optomechanical measurements are thus best suited\nfor investigating the interplay between the Barnett and\nEdH effects, such an application has been overlooked so\nfar.\nIn this Letter, we propose a levitated optomechanical\nsetup to directly probe angular-momentum transfer be-\ntween a spin system and mechanical rotation via Gilbert\ndamping with high precision. We consider a small levi-\ntated particle and study its uniaxial rotation under mi-\ncrowave irradiation in a ferromagnetic resonance (FMR)arXiv:2202.02461v2 [cond-mat.mes-hall] 19 Sep 20232\nFIG. 1. (a) Schematic picture of the model. The magnetic\nnanosphere is optically levitated. The static magnetic field\nH0is applied in the zdirection and the circularly-polarized\nmicrowave field in the xyplane. They induce precession of\nthe magnetization Mand macroscopic rotation Ωabout the\nzdirection via the spin-rotation coupling. (b) Net angular\nmomentum gain f(Ωz). There are three steady-state solu-\ntions, Ωz,i(i= 1,2,3). For Ωz,1andΩz,3(red points), an\ninfinitesimal change in Ωzproduces a restoring torque, and\nthus, the solutions become stable. In contrast, an infinites-\nimal deviation subsequently grows for Ωz,2, and thus, that\nsolution is unstable.\nexperiment (see Fig. 1 (a)). We calculate the steady-\nstate rotation frequency by balancing energy injection\nfrom microwaves and energy loss from air resistance and\nconclude that the particle can be rotated with a high ro-\ntation frequency up to GHz order in a vacuum. We show\nthat the steady-state solution exhibits bifurcation when\nthe rotation frequency is comparable to the microwave\nfrequency, which is the requirement for the strong Bar-\nnett effect, in other words, coexistence of the Barnett and\nthe EdH effects. The present setup enables us to sensi-\ntively measure the g-factor in the spin-rotation coupling.\nOur result illustrates the usefulness of levitated optome-\nchanical techniques in the study of gyromagnetism.\nDynamics of gyromagnetic systems.— We consider a\nspherical ferromagnetic levitated particle of radius r, in\nwhich the motion of the center of mass is suppressed.\nFor simplicity, the particle is regarded as a rigid body\nwith a moment of inertia, I= 2mptcr2/5, where mptc\nis the particle mass. We assume that the magnetization\nof the particle, M, is initially directed in the z-direction\nby an external static magnetic field. We consider a ferro-\nmagnetic resonance (FMR) experiment in which external\nmicrowaves irradiate a sample particle [55]. In this ex-\nperiment, the angular momentum of the excited spins is\ntransferred via the Gilbert damping to the rigid-body ro-\ntation of the particle and therefore the particle starts to\nrotate around the z-axis. Its rotation frequency vector is\ndenoted as Ω= (0,0, Ωz).\nFor the steady state, the Hamiltonian in the rotating\nframe fixed to the particle is written as [28, 29, 56]\nH=−(µ0γH+gSRΩ)·ℏStot. (1)\nThe first term describes the Zeeman energy, where Stot\nis the total spin of the particle, µ0is the vacuum perme-ability, γ(<0) is the gyromagnetic ratio, and His the\nmagnetic field in the rotating frame,\nH= (hcos(ω−Ωz)t, hsin(ω−Ωz)t, H 0). (2)\nHere, H0is a static magnetic field and h(≪H0) and\nωare the amplitude and frequency of the microwaves,\nrespectively. The second term of the Hamiltonian (1)\ndescribes the spin-rotation coupling that explains the\nEinstein-de Haas effect [2] and the Barnett effect [1].\nNote that the g-factor for the spin-rotation coupling, gSR,\ngenerally deviates from one [54, 56].\nThe magnetization of the particle is given as M=\nℏγStot/V, where V= 4πr3/3 is the volume of the par-\nticle. The Landau–Lifshitz–Gilbert (LLG) equation is\ngiven as\n˙M=M×(µ0γH+gSRΩ) +α\nM0M×˙M,(3)\nwhere M0=|M|,˙M= dM/dt, and αis the Gilbert\ndamping constant. The steady-state solution of the LLG\nequation can be obtained by assuming ˙Mz= 0 and\n(Mx, My) =M(cos [( ω−Ωz)t+ϕ],sin [(ω−Ωz)t+ϕ])\nand using M2\n0=M2+M2\nz. For this steady state, the\nzcomponent of the gain rate of the angular momentum\nfrom the spin system is given as\n[M×µ0γH]z=−α\nM0(ω−Ωz)M2. (4)\nThis coincides with the Gilbert damping term\nα/bracketleftig\nM×˙M/bracketrightig\nz/M0, indicating that the gain in an-\ngular momentum from the microwaves is lost through\nGilbert damping. This spin relaxation causes continuous\ntransfer of the angular momentum to the rigid-body\nangular momentum.\nMagnetically driven rigid-body rotation.— The equa-\ntion for the time evolution of the angular momentum in\nthe present system is given as [54]\nd\ndt/bracketleftbig\nIΩz+gSRℏStot\nz/bracketrightbig\n=Γin+Γair, (5)\nwhere Γin=µ0γℏ[Stot×H]zis the torque supplied from\nthe microwaves and Γairis the torque due to the air re-\nsistance (discussed later). When the Gilbert damping\nterm is included phenomenologically as in the last term\nof Eq. (3), a steady state, ˙Sz= 0, arises from the balance\nbetween the Gilbert damping and the gain in angular\nmomentum from the microwaves, Eq. (4). Then, Eq. (5)\nleads to the equation of motion for the steady-state ro-\ntation frequency Ωz.\nThe steady-state rotation of the particle receives\ntorque from the surrounding air. In this paper, we fo-\ncus on the molecular flow region in which λ≫rholds,\nwhere λis the mean free path, in order to realize high-\nfrequency rotation [43, 57–59]. We assume diffuse reflec-\ntion at the surface. Accordingly, the torque induced by3\nthe air resistance is calculated as [54]\nΓair=−8r4\n3/radicalbiggπmair\n2kBTΩzp, (6)\nwhere mairis the mass of the air molecules, kBis the\nBoltzmann constant, Tis temperature, and pis pressure.\nIn the following estimate, we assume that the particle is\nin the atmosphere, for which the average molecular mass\nismair= 4.78×10−26kg and take r= 1×10−6m and\nT= 273 K [60]. The mean free path satisfies p·λ=\nkBT/(√\n2πξ2), where ξis the diameter of an air molecule.\nSubstituting ξ= 3.76×10−10m, we have p·λ≃6.0×\n10−3N/m, and taking p≲100 Pa is sufficient to enter\nthe molecular flow region λ≫r.\nGyromagnetic bifurcation.— First, we set gSR= 1 in\norder to see the features of the steady-state rotation.\nThe Euler equation of a spherical particle is presented\nin Eq. (5). Defining the net angular momentum gain as\nf(Ωz)≡Γin(Ωz)+Γair(Ωz), the rotation frequency of the\nsteady state can be obtained by solving f(Ωz) = 0. Our\nestimate employs the parameters, M0= 1.557×105A/m,\nα= 6.7×10−5,H0= 2.6×105A/m, of a spin pumping\nexperiment for YIG [61].\nBefore showing our results, we should explain that\nboth stable and unstable solutions may appear in gen-\neral. Figure 1 (b) is a schematic graph of f(Ωz) and the\nsolutions of f(Ωz) = 0 when f(Ωz) has three solutions,\nΩz,i(i= 1,2,3). Two solutions, Ωz,1andΩz,3, are stable\nbecause an infinitesimal change in Ωzinduces a restoring\ntorque. On the other hand, Ωz,2is an unstable solution\nbecause no restoring torque works there. The emergence\nof three solutions highly depends on the microwave am-\nplitude hand the Gilbert damping parameter α, as will\nbe discussed later. Hereafter, we calculate the steady-\nstate rotation frequency as a function of the microwave\nfrequency ωand study how it changes as hvaries.\nFigures 2 (a)-(d) show the rotation frequency of the\nsteady state as a function of the microwave frequency for\nh= 4 A /m. The red and blue curves indicate the stable\nand unstable solutions, respectively. For a sufficiently\nhigh pressure, only one solution can be realized for ar-\nbitrary microwave frequencies (not shown in Fig. 2). As\nthe pressure is lowered, two stable solutions and one un-\nstable solution appear (Fig. 2 (a)). These new solutions\nproduced by bifurcation appear only near the resonant\nfrequency ω0=−µ0γH0≃57.52 GHz, while one stable\nsolution exists away from the resonant frequency. As the\npressure is lowered, the lower stable branch approaches\nthe upper one and they become connected to each other\nat a certain pressure, p1(Fig. 2 (b)). Below this crit-\nical pressure, the topology of the graph of Ωzchanges\n(Fig. 2 (c)), and a transition from the lower to the upper\nbranch becomes possible when the microwave frequency\nωapproaches ω0. For a sufficiently low pressure, the\nbifurcation disappears and only one stable solution ex-\nists for any microwave frequency (Fig. 2 (d)). In orderto see the characteristics of the branches, we show the\nmagnetization of the particle at the critical pressure in\nFig. 2 (e). The particle has a small (large) magnetization\nin the upper (lower) branch with a large (small) steady-\nstate rotation frequency. This means that the distribu-\ntion of the angular momentum between magnetization\nand rigid-body rotation is different in these two branches.\nFigure 2 (f) shows the steady-state solutions at ω=ω0\ninp-Ωzspace. The green area in the figure is the forbid-\nden region in which Mzbecomes an imaginary number.\nThe critical pressure is defined as the lowest pressure in\nthe lower branch in Fig. 2 (f) and is calculated as [54]\np1=4|γ|µ2\n0M0h2\nrαω2\n0/radicalbigg\nπkBT\n2mair. (7)\nIn the present estimate, the critical pressure at h=\n4 A/m isp1≃4.4×10−3Pa.\nFigures 2 (g)-(j) show the steady-state rotation fre-\nquency for h= 12 A /m. In this case, there is no bifur-\ncation and only one stable solution at any frequency and\nany pressure. Note that the graph of Ωzhas a cusp at\nω=ω0at a specific pressure p=p2. At this pressure, the\nmagnetization reaches zero at ω=ω0(Fig. 2 (k)). The\nbifurcation disappears because the unstable solutions are\ninside the forbidden region, as shown in Fig. 2 (l) which\nplots the steady-state solutions at ω=ω0on the p-Ωz\nplane. The pressure p=p2corresponds to the crossing\npoint of the two solutions, which is always on the bound-\nary of the forbidden region [54]. Finally, we should note\nthat for gSR= 1 the ferromagnetic resonance frequency\nis not affected by the rotation frequency. This is be-\ncause the decrease in microwave frequency in the rotat-\ning frame (see Eq. (2)) is completely compensated by the\ndecrease in the effective magnetic field due to the spin-\nrotation coupling (the Bernett effect) through the LLG\nequation (3).\nThus, whether the bifurcation occurs or not depends\non the pressure. The condition for a bifurcation to appear\nis obtained by solving p1< p 2atω=ω0ash/H 0<\nα/2 [54]. This indicates the bifurcation disappears at\nhigh microwave amplitudes, which is consistent with the\nresults shown in Fig. 2.\nNow, let us consider the case when gSRdeviates from\n1. Figure 3 shows the rotation frequency as a function of\nthe microwave frequency ωfor different values of gSRfor\np=p1andh= 4 A /m. The curves are strongly tilted\neven for a small deviation in gSRfrom 1. Furthermore,\neven when gSR−1 is small (see Fig. 3 (c),(d)), the two\nstable branches exist in a finite range of ω; therefore,\ntransitions from the lower branch to the upper branch\nalways occur as ωchanges, for example ω≃57.51 GHz\nin Fig. 3 (c). These changes originate from incomplete\ncompensation between the microwave frequency shift and\nthe spin-rotation coupling in the rotating frame. This\nbehavior can be utilized for measuring the effective g-\nfactor of the spin-rotation coupling. Note that if gSR−14\n(a) (b)\n(c) (d)(e)\n(g) (h)\n(i) (j)(k)\n(l)(f)\nFIG. 2. Steady-state rotation frequency Ωzfor various pres-\nsures with microwave amplitudes (a)–(d) 4 A /m and (g)–(j)\n12 A/m. (e)(k) The steady-state solutions of the magnetiza-\ntionMzthat correspond to Ωzin (b) and (h), respectively.\n(f)(l) Red and blue curves express the pdependence of the\nsteady-state solutions for the rotation frequency. The green\narea is the region in which the zcomponent of the magnetiza-\ntionMzbecomes imaginary. Thus, the physical solutions are\nthe red curves, and the solutions depicted by the blue curves\nare unstable.)\nis negative, the graph of Ωz(ω) is tilted in the opposite\ndirection.\nConclusion.— We examined angular momentum trans-\nfer between the spins and the mechanical angular mo-\nmentum of a levitated ferromagnetic particle driven by\nmicrowave irradiation in a vacuum. This setup is suit-\nable not only for making precise measurements but also\nfor very fast mechanical rotation, due to the absence of\na restoring torque, as shown in recent experiments on\nlevitated optomechanics [35]. We formulated the steady-\nstate rotation frequency using the LLG equation in com-\nbination with angular momentum conservation and es-\ntimated it using realistic experimental parameters. We\nfound a bifurcation phenomenon in the solutions for the\nsteady-state rotation frequency when the rotation fre-\nquency is very fast and comparable to the microwave\nfrequency. Our setup is a great candidate for satisfying\nthese conditions, which are essential for the strong Bar-\nnett effect, in other words, coexistence of the Barnett\nand the EdH effects. When the g-factor for the spin-\n(a) (b)\n(c) (d)\nFIG. 3. Steady-state solutions for the rotation frequency Ωz\nfor various gSR: (a) gSR= 1; (b) gSR= 1.0001; (c) gSR=\n1.001; (d) gSR= 1.01. The pressure and microwave amplitude\nare fixed: p=p1andh= 4 A /m. Red (blue) curves are stable\n(unstable). The scale of the horizontal axis is different in each\nfigure.\nrotation coupling, gSR, is unity, a transition in rotation\nfrequency is observed near the resonant frequency due\nto transitions between the two branches. Even a slight\ndeviation in the g-factor gSRfrom unity separates the\ntwo branches more significantly and the transition in ro-\ntation frequency appears in a wider region of pressure.\nThis feature of a bifurcation sensitive to the value of gSR\ncan be used for accurate measurement for the effective g-\nfactor. Our proposal will provide a powerful way to inves-\ntigate angular momentum conversion from magnetism to\nmacroscopic motion in ferromagnets and will promote in-\nterdisciplinary research between two developing research\nfields, spintronics and optomechanics.\nWe thank Y. Ominato, H. Taya, and M. Hongo for\nhelpful comments. We acknowledge JSPS KAKENHI for\nGrants (No. JP20K03831, No. JP21K03414). MM is par-\ntially supported by the Priority Program of the Chinese\nAcademy of Sciences, Grant No. XDB28000000. D.O. is\nsupported by the President’s PhD Scholarships at Im-\nperial College London, by JSPS Overseas Research Fel-\nlowship, by the Institution of Engineering and Technol-\nogy (IET), and by Funda¸ c˜ ao para a Ciˆ encia e a Tec-\nnologia and Instituto de Telecomunica¸ c˜ oes under project\nUIDB/50008/2020. T.S. was supported by the Japan So-\nciety for the Promotion of Science through the Program\nfor Leading Graduate Schools (MERIT).\n[1] S. J. Barnett, “Magnetization by rotation,” Phys. Rev.\n6, 239 (1915).\n[2] A. Einstein and W. J. de Haas, “Experimental proof of\nthe existence of amp` ere’s molecular currents,” KNAW5\nproc. 18 I, 696 (1915).\n[3] G. G. Scott, “Review of gyromagnetic ratio experi-\nments,” Rev. Mod. Phys. 34, 102 (1962).\n[4] Kazuya Harii, Yong-Jun Seo, Yasumasa Tsutsumi, Hi-\nroyuki Chudo, Koichi Oyanagi, Mamoru Matsuo, Yuki\nShiomi, Takahito Ono, Sadamichi Maekawa, and Eiji\nSaitoh, “Spin seebeck mechanical force,” Nature commu-\nnications 10, 1–5 (2019).\n[5] K. Mori, M. G. Dunsmore, J. E. Losby, D. M. Jenson,\nM. Belov, and M. R. Freeman, “Einstein–de Haas effect\nat radio frequencies in and near magnetic equilibrium,”\nPhys. Rev. B 102, 054415 (2020).\n[6] M. Matsuo, J. Ieda, E. Saitoh, and S. Maekawa, “Effects\nof mechanical rotation on spin currents,” Phys. Rev. Lett.\n106, 076601 (2011).\n[7] Mamoru Matsuo, Jun’ichi Ieda, Kazuya Harii, Eiji\nSaitoh, and Sadamichi Maekawa, “Mechanical genera-\ntion of spin current by spin-rotation coupling,” Physical\nReview B 87, 180402(R) (2013).\n[8] Atsufumi Hirohata, Yuji Baba, Benedict A Murphy,\nBenny Ng, Yunqi Yao, Kazuki Nagao, and Jun-young\nKim, “Magneto-optical detection of spin accumulation\nunder the influence of mechanical rotation,” Scientific re-\nports 8, 1–8 (2018).\n[9] R. Takahashi, M. Matsuo, M. Ono, K. Harii, H. Chudo,\nS. Okayasu, J. Ieda, S. Takahashi, S. Maekawa, and\nE. Saitoh, “Spin hydrodynamic generation,” Nat. Phys.\n12, 52 (2016).\n[10] R. Takahashi, H. Chudo, M. Matsuo, K. Harii,\nY. Ohnuma, S. Maekawa, and E. Saitoh, “Giant spin\nhydrodynamic generation in laminar flow,” Nat. Com-\nmun. 11, 3009 (2020).\n[11] Hamid Tabaei Kazerooni, Alexander Thieme, J¨ org Schu-\nmacher, and Christian Cierpka, “Electron spin-vorticity\ncoupling in pipe flows at low and high reynolds number,”\nPhys. Rev. Applied 14, 014002 (2020).\n[12] Hamid Tabaei Kazerooni, Georgy Zinchenko, J¨ org Schu-\nmacher, and Christian Cierpka, “Electrical voltage by\nelectron spin-vorticity coupling in laminar ducts,” Phys.\nRev. Fluids 6, 043703 (2021).\n[13] D. Kobayashi, T. Yoshikawa, M. Matsuo, R. Iguchi,\nS. Maekawa, E. Saitoh, and Y. Nozaki, “Spin current\ngeneration using a surface acoustic wave generated via\nspin-rotation coupling,” Phys. Rev. Lett. 119, 077202\n(2017).\n[14] Yuki Kurimune, Mamoru Matsuo, Sadamichi Maekawa,\nand Yukio Nozaki, “Highly nonlinear frequency-\ndependent spin-wave resonance excited via spin-vorticity\ncoupling,” Physical Review B 102, 174413 (2020).\n[15] Shoma Tateno, Genki Okano, Mamoru Matsuo, and\nYukio Nozaki, “Electrical evaluation of the alternating\nspin current generated via spin-vorticity coupling,” Phys-\nical Review B 102, 104406 (2020).\n[16] Shoma Tateno, Yuki Kurimune, Mamoru Matsuo,\nKazuto Yamanoi, and Yukio Nozaki, “Einstein–de haas\nphase shifts in surface acoustic waves,” Physical Review\nB104, L020404 (2021).\n[17] The STAR Collaboration, “Global Λhyperon polariza-\ntion in nuclear collisions,” Nature 548, 62 (2017).\n[18] Jaroslav Adam, L Adamczyk, JR Adams, James K Ad-\nkins, G Agakishiev, MM Aggarwal, Z Ahammed, NN Aji-\ntanand, I Alekseev, DM Anderson, et al. , “Global polar-\nization of λhyperons in au+ au collisions at s n n= 200\ngev,” Physical Review C 98, 014910 (2018).[19] Jaroslav Adam, L Adamczyk, JR Adams, JK Adkins,\nG Agakishiev, MM Aggarwal, Z Ahammed, I Alekseev,\nDM Anderson, R Aoyama, et al. , “Polarization of λ(λ¯)\nhyperons along the beam direction in au+ au collisions\nat s n n= 200 gev,” Physical review letters 123, 132301\n(2019).\n[20] Shreyasi Acharya, Dagmar Adamov´ a, Alexander Adler,\nJonatan Adolfsson, Madan M Aggarwal, G Aglieri\nRinella, Michelangelo Agnello, Nikita Agrawal, Zubayer\nAhammed, Shakeel Ahmad, et al. , “Evidence of spin-\norbital angular momentum interactions in relativistic\nheavy-ion collisions,” Physical review letters 125, 012301\n(2020).\n[21] Jaroslav Adam, L Adamczyk, JR Adams, JK Adkins,\nG Agakishiev, MM Aggarwal, Z Ahammed, I Alekseev,\nDM Anderson, A Aparin, et al. , “Global polarization of ξ\nandωhyperons in au+ au collisions at s n n= 200 gev,”\nPhysical review letters 126, 162301 (2021).\n[22] Thomas M Wallis, John Moreland, and Pavel Kabos,\n“Einstein–de Haas effect in a NiFe film deposited on a\nmicrocantilever,” Appl. Phys. Lett. 89, 122502 (2006).\n[23] G. Zolfagharkhani, A. Gaidarzhy, P. Degiovanni, S. Ket-\ntemann, P. Fulde, and P. Mohanty, “Nanomechanical\ndetection of itinerant electron spin flip,” Nat. Nanotech-\nnol.3, 720–723 (2008).\n[24] Masaki Imai, Yudai Ogata, Hiroyuki Chudo, Masao\nOno, Kazuya Harii, Mamoru Matsuo, Yuichi Ohnuma,\nSadamichi Maekawa, and Eiji Saitoh, “Observation of\ngyromagnetic reversal,” Applied Physics Letters 113,\n052402 (2018).\n[25] Masaki Imai, Hiroyuki Chudo, Masao Ono, Kazuya Harii,\nMamoru Matsuo, Yuichi Ohnuma, Sadamichi Maekawa,\nand Eiji Saitoh, “Angular momentum compensation ma-\nnipulation to room temperature of the ferrimagnet ho3-\nx dy x fe5o12 detected by the barnett effect,” Applied\nPhysics Letters 114, 162402 (2019).\n[26] C. Dornes, Y. Acremann, M. Savoini, M. Kubli, M. J.\nNeugebauer, E. Abreu, L. Huber, G. Lantz, C. A. F.\nVaz, H. Lemke, E. M. Bothschafter, M Porer, V. Espos-\nito, L. Rettig, M. Buzzi, A. Alberca, Y. W. Windsor,\nP. Beaud, U. Staub, Diling Zhu, Sanghoon Song, J. M.\nGlownia, and S. L. Johnson, “The ultrafast Einstein–de\nHaas effect,” Nature 565, 209–212 (2019).\n[27] W Izumida, R Okuyama, K Sato, T Kato, and M Mat-\nsuo, “Einstein–de haas nanorotor,” Physical Review Let-\nters128, 017701 (2022).\n[28] Friedrich W Hehl and Wei-Tou Ni, “Inertial effects of a\ndirac particle,” Physical Review D 42, 2045 (1990).\n[29] J¨ urg Fr¨ ohlich and Urban M. Studer, “Gauge invariance\nand current algebra in nonrelativistic many-body the-\nory,” Rev. Mod. Phys. 65, 733–802 (1993).\n[30] Arthur Ashkin, “Acceleration and trapping of particles\nby radiation pressure,” Physical Review Letters 24, 156\n(1970).\n[31] Arthur Ashkin and JM Dziedzic, “Optical levitation\nin high vacuum,” Applied Physics Letters 28, 333–335\n(1976).\n[32] Arthur Ashkin, James M Dziedzic, JE Bjorkholm, and\nSteven Chu, “Observation of a single-beam gradient force\noptical trap for dielectric particles,” Optics Letters 11,\n288–290 (1986).\n[33] Steven Chu, L. Hollberg, J. E. Bjorkholm, Alex Cable,\nand A. Ashkin, “Three-dimensional viscous confinement\nand cooling of atoms by resonance radiation pressure,”6\nPhys. Rev. Lett. 55, 48–51 (1985).\n[34] Vladan Vuleti´ c and Steven Chu, “Laser cooling of atoms,\nions, or molecules by coherent scattering,” Physical Re-\nview Letters 84, 3787 (2000).\n[35] C Gonzalez-Ballestero, M Aspelmeyer, L Novotny,\nR Quidant, and O Romero-Isart, “Levitodynamics: Lev-\nitation and control of microscopic objects in vacuum,”\nScience 374, eabg3027 (2021).\n[36] Tongcang Li, Simon Kheifets, and Mark G Raizen, “Mil-\nlikelvin cooling of an optically trapped microsphere in\nvacuum,” Nature Physics 7, 527–530 (2011).\n[37] Jan Gieseler, Bradley Deutsch, Romain Quidant, and\nLukas Novotny, “Subkelvin parametric feedback cooling\nof a laser-trapped nanoparticle,” Physical Review Letters\n109, 103603 (2012).\n[38] Vijay Jain, Jan Gieseler, Clemens Moritz, Christoph Del-\nlago, Romain Quidant, and Lukas Novotny, “Direct mea-\nsurement of photon recoil from a levitated nanoparticle,”\nPhysical Review Letters 116, 243601 (2016).\n[39] Darrick E Chang, CA Regal, SB Papp, DJ Wilson, J Ye,\nO Painter, H Jeff Kimble, and P Zoller, “Cavity opto-\nmechanics using an optically levitated nanosphere,” Pro-\nceedings of the National Academy of Sciences 107, 1005–\n1010 (2010).\n[40] J. Millen, P. Z. G. Fonseca, T. Mavrogordatos, T. S.\nMonteiro, and P. F. Barker, “Cavity cooling a single\ncharged levitated nanosphere,” Phys. Rev. Lett. 114,\n123602 (2015).\n[41] Felix Tebbenjohanns, Martin Frimmer, Vijay Jain, Do-\nminik Windey, and Lukas Novotny, “Motional sideband\nasymmetry of a nanoparticle optically levitated in free\nspace,” Physical Review Letters 124, 013603 (2020).\n[42] Ren´ e Reimann, Michael Doderer, Erik Hebestreit,\nRozenn Diehl, Martin Frimmer, Dominik Windey, Fe-\nlix Tebbenjohanns, and Lukas Novotny, “Ghz rotation\nof an optically trapped nanoparticle in vacuum,” Phys.\nRev. Lett. 121, 033602 (2018).\n[43] Jonghoon Ahn, Zhujing Xu, Jaehoon Bang, Yu-Hao\nDeng, Thai M. Hoang, Qinkai Han, Ren-Min Ma, and\nTongcang Li, “Optically levitated nanodumbbell torsion\nbalance and ghz nanomechanical rotor,” Phys. Rev. Lett.\n121, 033603 (2018).\n[44] J C¯ ımurs and A C¯ ebers, “Three-dimensional dynamics of\na particle with a finite energy of magnetic anisotropy in\na rotating magnetic field,” Physical Review E 88, 062315\n(2013).\n[45] J C¯ ımurs and A C¯ ebers, “Dynamics of anisotropic super-\nparamagnetic particles in a precessing magnetic field,”\nPhysical Review E 87, 062318 (2013).\n[46] Klaus D Usadel, “Dynamics of magnetic nanoparticles\nin a viscous fluid driven by rotating magnetic fields,”\nPhysical Review B 95, 104430 (2017).\n[47] Klaus D Usadel and Clemens Usadel, “Dynamics of mag-\nnetic single domain particles embedded in a viscous liq-\nuid,” Journal of Applied Physics 118, 234303 (2015).\n[48] NA Usov and B Ya Liubimov, “Magnetic nanoparticle\nmotion in external magnetic field,” Journal of Magnetismand Magnetic Materials 385, 339–346 (2015).\n[49] NA Usov and B Ya Liubimov, “Dynamics of magnetic\nnanoparticle in a viscous liquid: Application to magnetic\nnanoparticle hyperthermia,” Journal of Applied Physics\n112, 023901 (2012).\n[50] Hedyeh Keshtgar, Simon Streib, Akashdeep Kamra,\nYaroslav M Blanter, and Gerrit EW Bauer, “Magne-\ntomechanical coupling and ferromagnetic resonance in\nmagnetic nanoparticles,” Physical Review B 95, 134447\n(2017).\n[51] TV Lyutyy, Oleksandr Mykolaiovych Hryshko, and\nM Yu Yakovenko, “Uniform and nonuniform precession\nof a nanoparticle with finite anisotropy in a liquid: Op-\nportunities and limitations for magnetic fluid hyperther-\nmia,” Journal of Magnetism and Magnetic Materials 473,\n198–204 (2019).\n[52] TV Lyutyy, Stanislav I Denisov, and Peter H¨ anggi,\n“Dissipation-induced rotation of suspended ferromag-\nnetic nanoparticles,” Physical Review B 100, 134403\n(2019).\n[53] P. Huillery, T. Delord, L. Nicolas, M. Van Den Bossche,\nM. Perdriat, and G. H´ etet, “Spin mechanics with levitat-\ning ferromagnetic particles,” Phys. Rev. B 101, 134415\n(2020).\n[54] See Supplemental Material at\nhttp://link.aps.org/supplemental/10.1103/PhysRevLett.***.******\nfor experimental implementation, detailed calculations of\nthe air resistance, the angular-momentum conservation\nlaw, the critical pressure, effect of the Gilbert damping,\ng-factor for the spin-rotation coupling, and stability\nanalysis of the rotation.\n[55] Charles Kittel, “On the theory of ferromagnetic reso-\nnance absorption,” Phys. Rev. 73, 155–161 (1948).\n[56] Mamoru Matsuo, Jun’ichi Ieda, and Sadamichi\nMaekawa, “Renormalization of spin-rotation coupling,”\nPhys. Rev. B 87, 115301 (2013).\n[57] Paul S. Epstein, “On the resistance experienced by\nspheres in their motion through gases,” Phys. Rev. 23,\n710–733 (1924).\n[58] A. Roth, Vacuum Technology (North Holland, 1990).\n[59] James Corson, George W. Mulholland, and Michael R.\nZachariah, “Calculating the rotational friction coefficient\nof fractal aerosol particles in the transition regime using\nextended kirkwood-riseman theory,” Phys. Rev. E 96,\n013110 (2017).\n[60] In our estimate, a fluctuation of rotation velocity, which\nis dominantly induced by thermal fluctuation of air re-\nsistance torque, becomes much smaller than its average.\nTherefore, the noise of rotation frequency does not affect\nits precious measurement.\n[61] Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida,\nM. Mizuguchi, H. Umezawa, H. Kawai, K. Ando,\nK. Takanashi, S. Maekawa, and E. Saitoh, “Transmis-\nsion of electrical signals by spin-wave interconversion in\na magnetic insulator,” Nature 464, 262–266 (2010).Supplementary Information:\nGyromagnetic bifurcation in a levitated ferromagnetic particle\nT. Sato,1T. Kato,1Daigo Oue,2, 3, 4and M. Matsuo2, 5, 6, 7\n1Institute for Solid State Physics, University of Tokyo, Kashiwa 277-8581, Japan\n2Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing, 100190, China.\n3The Blackett Laboratory, Imperial College London, London SW7 2AZ, United Kingdom\n4Instituto Superior T´ ecnico, University of Lisbon, 1049-001 Lisboa, Portugal\n5CAS Center for Excellence in Topological Quantum Computation,\nUniversity of Chinese Academy of Sciences, Beijing 100190, China\n6Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, 319-1195, Japan\n7RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan\n(Dated: September 20, 2023)\nMEASUREMENT OF THE ROTATION\nFREQUENCY\nThe fast rotation of an optically levitated particle has\nbeen observed in a number of previous works [1–4]. For\nexample, Ref. [2] reported a GHz rotation of a single\n100 nm silica particle in vacuum in pressure range from\n10−3Pa to 10 Pa. In Ref. [2], the authors analyzed the\nfrequency shift of the detection beam which goes through\nthe particle. Because this shift is independent of the ro-\ntation detection mechanism, it has been utilized for accu-\nrate measurement of the rotation frequency of the parti-\ncle. Though these experiments are mainly performed for\nnon-magnetic particles, levitation of ferromagnetic parti-\ncles has also been realized recently. For example, Ref. [5]\nachieved the levitation of a spherical iron particle having\na diameter around 1 µm by a Paul trap method. There-\nfore, we expect that both levitation of ferromagnetic par-\nticles and measurement of their rotation frequency that\nare assumed in our work are possible within the present\nexperimental techniques. We note that our theoretical\ndescription based on the LLG equation and the angular-\nmomentum conservation does not depends on the kind\nof magnetic materials and is applicable to various ferro-\nmagnetic particles.\nAIR RESISTANCE IN VACUUM\nWe formulate the torque by the air resistance in vac-\nuum referring to Refs. [3, 6–8]. We assume the condition\nofλ≫r, where λis the mean free path of the air and\nris the radius of the particle. We further assume the\ndiffuse reflection at the surface. Then, the total torque\nof the air resistance for the particle is given as [9]\nΓair=−2πr4Ωz\n3ρ¯c, (1)\nwhere ρand ¯care the density and the mean velocity of\nmolecules, respectively. From the Maxwell-Boltzmann\ndistribution for an ideal gas, these quantities are givenp mean free path λ\n10−3Pa 6 m\n10−2Pa 60 cm\n0.1 Pa 6 cm\n1 Pa 6 mm\n10 Pa 0 .6 mm\nTABLE I. The pressure dependence of the mean free path for\nthe air molecules.\nas\nρ=nmair=mairp\nkBT,¯c=2√π/radicalbigg\n2kBT\nmair, (2)\nwhere mairis the molecule mass, nis the number of\nmolecules per unit volume, Tis the temperature, pis\nthe pressure, and kBis the Boltzmann constant. Substi-\ntuting Eq. (2) into Eq. (1), we obtain\nΓair=−8r4\n3/radicalbiggπmair\n2kBTpΩz≡ −βΩz. (3)\nWe assume the mass of air molecules, m= 4.78×\n10−26kg. We also take r= 10−6m and T= 273 K.\nTo check the condition λ≫r, we need to evaluate the\nmean free path, which is given in Ref. 7 as\nλ=kBT√\n2πξ2p(4)\nwhere ξis the diameter of an air molecule. Taking ξ=\n3.76×10−10m [10] and T= 273 K, the mean free path\nof the air satisfies\np·λ≃6.0×10−3N/m. (5)\nThe mean free path for several typical pressures is given\nin Table I. For r= 10−6m, the condition λ≫ris well\nsatisfied when p≲100 Pa.arXiv:2202.02461v2 [cond-mat.mes-hall] 19 Sep 20232\nDERIVATION OF THE ANGULAR-MOMENTUM\nCONSERVATION LAW\nWe derive the angular-momentum conservation law of\nthe system. For this purpose, it is convenient to start\nfrom the Lagrangian in the rotating frame given as\nL=1\n2I˙φ2+ℏStot˙ϕs(cosθs−1)− H\n=1\n2I˙φ2+ℏStot˙ϕs(cosθs−1) +µ0γH·S+gSRΩ·S\n=1\n2I˙φ2+ℏStot˙ϕs(cosθs−1)\n+ℏStotµ0γ{hsinθs[cos(ωt−φ) cosϕs\n+ sin( ωt−φ) sinϕs] +H0cosθs}\n+gSRℏStotcosθs˙φ, (6)\nwhere Stot=Stot(sinθscosϕs,sinθssinϕs,cosθs),φis\na rotation angle whose time derivative corresponds to\nΩ= (0,0,˙φ), and we have imposed φ(t= 0) = 0 as\nthe initial condition. From the Euler-Lagrange equation\nwith respect to φ,d\ndt∂L\n∂˙φ=∂L\n∂φ, we obtain\nd\ndt/bracketleftbig\nI˙φ+gSRℏStot\nz/bracketrightbig\n=ℏStotµ0γh[sin(ωt−φ) cosϕs−cos(ωt−φ) sinϕs]\n=µ0γℏ/bracketleftbig\nStot×H/bracketrightbig\nz≡Γin, (7)\nwhere Γin=µ0γℏ[Stot×H]zis the torque supplied from\nthe microwave. By introducing the dissipative function\nto describe the air resistance, an additional torque Γair\nis added to the right-hand side of the above equation.\nBy considering the steady state, ˙ φ=Ωz, and using φ=/integraltextt\n0˙φdt=Ωzt, we obtain Eq. (7) in the main text.\nSTEADY-STATE EULER EQUATION\nThe dynamics of the system is determined by the LLG\nequation\n˙M=M×(µ0γH+gSRΩ) +α\nMM×˙M,(8)\nand the Euler equation\nI˙Ωz+gSRℏ˙Stot\nz=Γin+Γair, (9)\nwhere we have assumed Ω= (0,0, Ωz). We further as-\nsume the xandycomponents of Mas\nMx=Mcos [(ω−Ωz)t+ϕ], (10)\nMy=Msin [(ω−Ωz)t+ϕ],\nwhere M2=M2\n0−M2\nz. By taking\n˙Mz=˙Ωz= 0, (11)the steady-state solution is obtained as\ntanϕ=αMz\nM0(ω−Ωz)\nω−ω0+∆gΩ z, (12)\nM=µ0γMzh/radicalbig\n(ω−ω0+∆gΩ z)2+ (α(ω−Ωz)Mz/M0)2.\n(13)\nwhere ω0=−µ0γH0and we defined ∆g=gSR−1. Using\nM2+M2\nz=M2\n0, the equation for M2\nzcan be derived. By\nsolving this equation, we obtain\nM2\nz=1\n2α2(ω−Ωz)2/M2\n0/bracketleftig\n−A+/radicalbig\nA2+B/bracketrightig\n, (14)\nA=µ2\n0γ2h2+ (ω−ω0+∆gΩ z)2−α2(ω−Ωz)2,\n(15)\nB= 4α2(ω−Ωz)2(ω−ω0+∆gΩ z)2. (16)\nHere, we have chosen the solution in which the sign in\nfront of the square root is plus, because Mzhas to be\nreal. Since Γinis written as\nΓin=µ0γℏ/bracketleftbig\nStot×H/bracketrightbig\nz\n=−αV\nM0γ(ω−Ωz)(M2\n0−M2\nz), (17)\nthe steady-state Euler equation (9) becomes\n−αV\nM0γ(ω−Ωz)(M2\n0−M2\nz)−βΩz= 0, (18)\nwhere βis the constant defined in Eq. (3). Substituting\nEq. (14) into Eq. (18), we obtain the cubic equation for\nΩzas\na3Ω3\nz+a2Ω2\nz+a1Ωz+a0= 0, (19)\nwhere\na3=−/parenleftbigg2βα\nM0/parenrightbigg2\n+4βα\nM0V\nγ(∆g2+α2), (20)\na2=/parenleftbigg2βα\nM0/parenrightbigg2\nω+8βα\nM0V\nγ(∆g(ω−ω0)−α2ω),(21)\na1=4βα\nM0V\nγ(µ2\n0γ2h2+ (ω−ω0)2+α2ω2)\n−4V2\nγ2µ2\n0γ2h2α2, (22)\na0= 4V2\nγ2µ2\n0γ2h2α2ω. (23)\nThe solution of Eq. (19) has to satisfy the condition\nh(Ωz) =/radicalbig\nA(Ωz)2+B(Ωz)\n=µ2\n0γ2h2+ (ω−ω0+∆gΩ z)2+α2(ω−Ωz)2\n+2αβγ\nM0VΩz(ω−Ωz)≥0, (24)\nwhich appears when squaring Eq. (18). The solutions\nof Eq. (19) which don’t satisfy Eq. (24) correspond to\nnegative M2\nz, in other words imaginary Mz.3\nD⒩\nD⒪D⒩D⒪(a) (b)\nFIG. 1. The steady-state rotation frequency Ωzas a function\nof the pressure for ∆g= 0 and ω=ω0at (a) h= 4 A /m and\n(b)h= 12 A /m. Red (blue) solid curves express the steady-\nstate stable (unstable) solutions of rotation frequency. Green\narea is the region that the zcomponent of the magnetization\nMzbecomes imaginary. When solving Eq. (19), not only the\nred and blue solid curves but also red dotted curves appear.\nHowever, red dotted curves are prohibited since its zcompo-\nnent of the magnetization becomes imaginary.\nBIFURCATION AND CRITICAL PRESSURE\nIn this section, we describe the condition for bifurca-\ntion and detailed calculation on the critical pressure for\ngSR= 1 and ω=ω0given in the main text. We show the\nreal solutions of Ωzas a function of the pressure pby the\nsolid lines in Fig. 1. The red (blue) part of the solid line\nindicates the stable (unstable) solution. Fig. 1 (a) and (b)\ncorrespond to the cases of h= 4 A /m and h= 12 A /m,\nrespectively. We also show the forbidden region, in which\nthe condition (24) does not holds, by the green hatch in\nFig. 1. The dashed lines in the hatch indicate the un-\nphysical solutions, which give imaginary values for Mz.\nBy comparison between Fig. 1 (a) and (b), we can\nsee that the behavior of the steady-state solutions is dis-\ntinct for h= 4 A /m and h= 12 A /m; one unstable so-\nlution exists for weak amplitude of microwave, while it\ndisappears for large amplitude of microwave. This dif-\nference is analyzed by the relative position between the\nbifurcation point D1and the crossing point D2shown in\nFig. 1. These two points correspond to the multiple roots\nof Eq. (19) for ∆g= 0 and ω=ω0. We define x,y, and\nzas\nx=2βα\nM0, y=−V\nγα2, z=−V\nγµ2\n0γ2h21\nω2\n0.(25)\nWe note that x,y, and zare all positive since γ < 0.\nThen, the cubic equation given in Eq. (19) is rewritten\nfor∆g= 0 and ω=ω0as\nΩ3\nz(x+ 2y)x−Ω2\nz(x+ 4y)xω0 (26)\n+Ωz(4yz+ 2xz+ 2xy)ω2\n0−4yzω3\n0= 0.\nThe discriminant of Eq.(26) is given as\nD= 4ω4\n0x(x−8z)(x2y−x2z−4xyz−4y2z)2.(27)By solving D= 0, the condition for the multiple roots is\ngiven as x=x1, x2where\nx1= 8z, x 2=2y(z+√yz)\ny−z. (28)\nWe note that the sign in front of the square root is chosen\nas a plus since xhave to be positive. The solutions, x=\nx1andx=x2, correspond to D1andD2, respectively.\nThe steady-state rotational frequencies are calculated for\nthese two solutions as\nΩz(D1) =1\n2ω0, (29)\nΩz(D2) =y−z\ny+√yzω0. (30)\nNext, we consider the boundary of the forbidden area\n(the green hatched area in Fig. 1). It is written with x,\ny, and zas\nh(Ωz) =−γ\nV/bracketleftig\n(x+y)Ω2\nz+ (−x−2y)ω0Ωz\n+(y+z)ω2\n0/bracketrightig\n= 0. (31)\nWe can check that the solution for D2given by x=\nx2and Eq. (30) satisfies this condition. This indicates\nthat the crossing point D2of the steady-state solution is\nalways on the boundary of the forbidden region in Fig. 1.\nWe note that the D2point disappears when y < z ; in this\ncase only one branch is observed just like the usual FMR.\nFory > z , the unstable solution (indicated by the blue\nline in Fig. 1 (a)) exists only when Ωz(D1)< Ω z(D2).\nThis condition is written as\nh\nH0<α\n2. (32)\nWhen the unstable solution exists, the D1point cor-\nresponds to the critical pressure p1at which the two\nbranches connect each other:\np1=−2M0\nγαr/radicalbigg\n2πkBT\nmair/parenleftbiggh\nH0/parenrightbigg2\n. (33)\nForh= 4 A /m, which correspond to the situation given\nin Fig. 3 (b) in main text, the critical pressure is evalu-\nated as p1≃4.4×10−3Pa.\nEFFECT OF THE GILBERT DAMPING\nWe discuss the effect of the Gilbert damping for the\nsteady-state rotation. We first show the rotation fre-\nquency Ωzas a function of the microwave frequency ω\nforgSR= 1,p= 5.0×10−3Pa, and h= 4 A /m in Fig. 2.\nAs the Gilbert damping αis increased from Fig. 2 (a) to\n(c), the bifurcation first occurs in the lower branch and\nsubsequently the upper branch is disconnected from the4\n(a)\n(c) (d)\n(b)\nFIG. 2. Steady real solutions of rotation frequency for each α\natgSR= 1, p= 5.0×10−3Pa and h= 4 A /m are described.\nαcr≃5.89×10−5determined from Eq. (33).\nlower one. Finally, the pair of the stable and unstable\nsolutions disappears after the bifurcation in the upper\nbranch and only one stable solution is left as shown in\n(d). The increase of αgives a similar effect as the de-\ncrease of the amplitude of the microwave, h.\nNext, we show the rotation frequency for gSR= 1.001,\np= 5.0×10−3Pa, and h= 4 A /m. If αis sufficiently\nsmall, there are two stable solutions (red lines) and one\nunstable one (a blue line) as shown in Fig. 3 (a) and\n(b), and the shape of Ωz(ω) is tilted as pointed out in\nthe main text. As αincreases, the critical pressure p1\nbecomes lower than the present pressure and the upper\nbranch is disconnected from the lower branch as shown\nin Fig. 3 (c). If αfurther increases, the upper branch\ndisappears and only one solution is realized as shown\nin Fig. 3 (d), for which the tilt of the graph of Ωz(ω)\nbecomes obscure because of large broadening of the reso-\nnant peak. We note that if the pressure pis much lower,\ntwo separated branches exist for any value of the Gilbert\ndamping αsince the pressure at the D2point approaches\nto∼4.7×10−3Pa.\nIn the main text, we chose α= 6.7×10−5from a\ntypical value for yttrium iron garnet (YIG). The tilting of\nthe ferromagnetic resonance (FMR) peak for Ωzcannot\nbe observed if αbecomes large as shown in Fig. 3 (d).\nHowever, the tilting of the FMR peak can be recovered\nby increasing the microwave amplitude hfor such a large\nvalue of the Gilbert damping.\n(a)\n(c) (d)\n(b)\nFIG. 3. Steady real solutions of rotational frequency Ωzfor\ndifferent αare described. gSR,p, and his fixed to 1 .001 and\n5.0×10−3Pa, and 4 A /m respectively.\ng-FACTOR FOR THE SPIN-ROTATION\nCOUPLING\nIn the main text, we have pointed out the possibility\nthat the g-factor for the spin-rotation coupling, gSR, de-\nviates from 1. Here, we briefly outline how gSRis shifted\nfrom 1 for ferromagnetism induced by localized electrons\nin 3dtransition-metal ions by considering the crystal field\nand the spin-orbit coupling following Ref. 11.\nWe consider a localized spin in an ion of 3 dtransition\nmetals whose Hamiltonian is given as\nH=Hintra+Hcry+HSO+HZ+HSR. (34)\nHere, HintraandHcrydescribe the intraatomic Coulomb\nand the crystal field, respectively. The last three terms\nare the Hamiltonians of the spin-orbit coupling, the Zee-\nman energy, and the spin-rotation coupling and are given\nas\nHSO=λL·S, (35)\nHZ=µBH·(L+ 2S), (36)\nHSR=Ω·(L+S), (37)\nrespectively. We assume H0=Hintra +Hcry≫\nHSO, HZ, HSR, which is satisfied for 3 dtransition-metal\nions. We assume that the ground state of H0is nonde-\ngenerate and express it as |G, M s⟩, where Msis the spin\nquantum number. Since H0does not mix spin and orbit\nstates, this eigenfunctions are written as the products,\n|G, M s⟩=|G⟩|Ms⟩. Using the perturbation theory, the\nfirst- and the second-order corrections to the energy for\nan orbitally nondegenerate ground state, E(1), E(2), are5\nderived as\nE(1)= 2µBH·S+Ω·S, (38)\nE(2)=−/summationdisplay\ni̸=G|⟨G|L·(λS+µBH+Ω)|i⟩|2\nE(0)\ni−E(0)\nG(39)\nwhere we have used the fact that ⟨G|L|G⟩vanish. When\nwe define\nΛµν=/summationdisplay\ni̸=G⟨G|Lµ|i⟩⟨i|Lν|G⟩\nE(0)\ni−E(0)\nG, (40)\nthe energy correction reduces to\nE(2)=−Λµν/bracketleftbig\nλ2SµSν+ (µBHµ+Ωµ)(µBHν+Ων)\n+2λSµ(µBHν+Ων)] (41)\nwhere the Einstein summation convention has been used.\nBy combining these results, the effective Hamiltonian be-\ncomes\nHeff=H0+gZ\nµνµBHµSν+gSR\nµνΩµSν−λ2ΛµνSµSν\n−Λµν(µBHµ+Ωµ)(µBHν+Ων), (42)\ngZ\nµν= 2δµν−2λΛµν, (43)\ngSR\nµν=δµν−2λΛµν. (44)\nwhere gZ\nµνandgSR\nµνareg-factor tensors for the Zeeman\nterm and the spin-rotation coupling, respectively. This\nresult indicates that the spin-orbit interaction changes\nnot only the g-factor of the Zeeman term but also that of\nthe spin-rotation coupling in the same way through Λµν.\nIn general, Λµνcan be diagonalized by taking a certain\nprinciple axis. If the crystal field has cubic symmetry,\nΛµνbecomes\nΛµν= 0 (µ̸=ν), Λxx=Λyy=Λzz. (45)\nThis means an isotropic constant shift of the g-factor. In\nthe main text, we have assumed an isotropic change of\ntheg-factor for simplicity.\nSTABILITY ANALYSIS OF THE ROTATION\nThrough our article, we have assumed that the in-plane\nrotation velocities, ΩxandΩy, are zero. We have checked\nnumerically that the in-plane rotation always decays even\nwhen the initial condition includes the in-plane rotation.\nThis indicates that the uniaxial rotation around the z\naxis is stable against perturbation to incline the rotation\naxis. In this section, we show the stability of the uniaxial\nrotation by solving Eqs. (48) and (56) within a simple\napproximation, which leads to the same result obtained\nby numerical calculation.\nIn order to discuss a general formalism of our model,\nwe have to ride on the body frame in which the in-plane\nrotation is also allowed. The coordinate transformationfrom the laboratory frame, ex,ey,ez, to the body frame,\nex′,ey′,ez′, can be implemented by using Euler angles,\n(φ, θ, ψ ). The transformation matrix R ( Ab= RA,\nwhere Ais a vector in the lab frame, and Abis that\nin the body frame) is given as\nR =\ncosψsinψ0\n−sinψcosψ0\n0 0 1\n·\n1 0 0\n0 cos θsinθ\n0−sinθcosθ\n (46)\n·\ncosφsinφ0\n−sinφcosφ0\n0 0 1\n.\nThe rotation frequency Ωis expressed by Euler angles\nas\n\n\nΩx′= ˙φsinθsinψ+˙θcosψ,\nΩy′= ˙φsinθcosψ−˙θsinψ,\nΩz′= ˙φcosθ+˙ψ.(47)\nWhen θ=φ= 0,˙θ= ˙φ= 0, the body frame rotates only\naround zaxis, and it becomes the same as the model of\nour article. For stability analysis, it is sufficient to con-\nsider states of θ≪1. The dynamics of the magnetization\nin the body frame, Mb, is determined from a modified\nversion of Eq. (8):\n˙Mb=Mb×(µ0γHb+Ωb) +α\nMMb×˙Mb, (48)\nwhere gSRis set as unity for simplicity and Hbis defined\nas\nHb= RH= R\nhcosωt\nhsinωt\nH0\n (49)\n≃\nhcos(ωt−φ−ψ)\nhsin(ωt−φ−ψ)\nH0\n+θ\nH0sinψ\nH0cosψ\n−hsin(ωt−φ)\n.\nBy expanding Mbwith respect to θas\nMb=M+θM1+O(θ2), (50)\nwe can derive equations from the zeroth- and first-order\nofθ. The zeroth-order equation leads to the equation of\nMgiven in the main text and therefore the steady-state\nsolution of Mis given as\nM≡\nMx′\nMy′\nMz′\n=\nMcos[(ω−Ωz′)t+ϕ]\nMsin[(ω−Ωz′)t+ϕ]\nMz\n. (51)\nHere, we used Ωz′= ˙φcosθ+˙ψ≃˙φ+˙ψand as-\nsumed that ˙ φand ˙ψare almost constant. On the\nother hand, from the equation of the first order of θ,\nM1≡(M1x′, M1y′, M1z′) obeys6\n\n∂tα\nM0Mz′∂t−µ0γH0−Ωz′−α\nM0My′∂t+µ0γHy′+α\nM0˙My′\n−α\nM0Mz′∂t+µ0γH0+Ωz′ ∂tα\nM0∂t−µ0γHx′−α\nM0˙Mx′\nα\nM0My′∂t−µ0γHy′−α\nM0˙My′−α\nM0Mx′∂t+µ0γHx′−α\nM0˙Mx′ ∂t\n\nM1x′\nM1y′\nM1z′\n\n=−µ0γ\nH0Mz′cosψ+hsin(ωt−φ)My′\n−H0Mz′sinψ−hsin(ωt−φ)Mx′\nH0(sinψMy′−cosψMy′)\n, (52)\nwhere ( Hx′, Hy′)≡h(cos(ω−Ωz′)t,sin(ω−Ωz′)t). Ex- tracting upper two equations, and assuming h≪H0,\nthese equations are reduced to\n/parenleftig\n∂t−i˙ψ+i/parenleftig\n−α\nM0Mz′(∂t−i˙ψ) +µ0γH0+Ωz′/parenrightig\niα\nM0M+∂t−iµ0γH++α\nM0(ω−˙φ−˙ψ)M+/parenrightig/parenleftigg\nM1\nM1z′/parenrightigg\n=−µ0γH0Mz′, (53)\nwhere M1≡(M1x′+iM1y′)eiψ,M+≡Mei(ωt−φ+ϕ)and\nH+≡(Hx′+iHy′)eiψ=hei(ωt−φ).M+andH+oscil-\nlate with the microwave frequency ωbecause ˙ φis much\nsmaller than ω. Since the dynamics of M1is roughly\ndominated by slowly oscillating terms, we can drop these\nfast oscillating terms. Technically speaking, M+and\nH+becomes zero by being integrated out from t= 0\ntot= 2π/ω. Then, Eq. (53) becomes\n/parenleftbigg\n1−iα\nM0Mz′/parenrightbigg\n∂tM1 (54)\n+/bracketleftbigg\ni(µ0γH0+Ωz′−˙ψ)−α\nM0Mz′˙ψ/bracketrightbigg\nM1=−µ0γMz′,\nand the steady-state solution of this is obtained as\nM1=−µ0γH0Mz′\ni(µ0γH0+Ωz′−˙ψ)−α\nM0Mz′˙ψ. (55)\nEuler equation of the particle is written as\n\n\nI1˙Ωx′+∆IΩ z′Ωy′= [Γg+Γoptical +Γair]x′,\nI1˙Ωy′−∆IΩ z′Ωx′= [Γg+Γoptical +Γair]y′,\nI2˙Ωz′= [Γg+Γoptical +Γair]z′,(56)\nwhere we assumed that the particle is symmetrical top\nandI1(I2) is a principle moment of inertia around x′, y′\n(z′) axis and ∆I≡I2−I1. Here, Γgis a torque generated\nfrom Gilbert damping given as\nΓg=−V\nγα\nM0Mb×˙Mb, (57)Γoptical is a torque from optical trap, and Γair=−βΩis\na torque from air resistance. Using Eq. (47), equation of\nmotion for θ,φandψis written as\nI1¨θ+I1˙φ˙ψsinθ+∆IΩ z′˙φsinθ\n= [Γg+Γoptical +Γair]x′cosψ\n−[Γg+Γoptical +Γair]y′sinψ, (58)\nI1( ¨φsinθ+ ˙φ˙θcosθ−˙θ˙ψ)−∆IΩ z′˙θ\n= [Γg+Γoptical +Γair]x′sinψ\n+ [Γg+Γoptical +Γair]y′cosψ, (59)\nI2(¨ψ+ ¨φcosθ−˙φ˙θsinθ) = [Γg+Γoptical +Γair]z′.\n(60)\nNeglecting the fast oscillating term, the first term in the\nbracket of the right hand side is approximated as\nΓg,x′cosψ−Γg,y′sinψ≃ −V\nγα\nM0θ˙ψMz′Re(M1),(61)\nΓg,x′sinψ+Γg,y′cosψ≃ −V\nγα\nM0θ˙ψMz′Im(M1),(62)\nΓg,z′≃ −V\nγα\nM0(ω−Ωz′)(M2\n0−M2\nz′). (63)\nThe torque from optical trap, Γoptical , which usually\nworks to restore the axis of inertia to zaxis, is written\nas\nΓoptical ,x′cosψ−Γoptical ,y′sinψ=−ξθ, (64)\nΓoptical ,x′sinψ+Γoptical ,y′cosψ= 0, (65)\nΓoptical ,z′= 0, (66)7\nwhere ξis a coefficient of the optical torque. In addition,\nwe take the spherical particle limit ∆I=I2−I1→0\nas considered in the main text. By these assumptions,\nEq. (58) and (59) are simplified as\nI1¨θ+β˙θ+/parenleftbigg\nI1˙φ˙ψ+ξ+V\nγα\nM0˙ψMz′Re(M1)/parenrightbigg\nθ≃0,\n(67)\nI1θ¨φ+ (I1˙θ+βθ) ˙φ−I1˙θ˙ψ+V\nγα\nM0θ˙ψMz′Im(M1)≃0,\n(68)\nI1(¨ψ+ ¨φ)≃ −βΩz′−V\nγα\nM0(ω−Ωz′)(M2\n0−M2\nz′).\n(69)\nIn our setup, we estimate I1≃9×10−27N·m·s2,β≃\n5×10−29N·m·s and\nV\nγα\nM0Mz′Im(M1)≃V\nγαM0ω0−˙φ\nω0\n∼(ω0−˙φ)×(−4.3×10−39) N·m·s,(70)\nV\nγα\nM0Mz′Re(M1)≃ −V\nγα2M0˙ψ\nω0\n∼˙ψ×2.9×10−43N·m·s. (71)\nIn addition, since the optical force is typically order of\nsub-pico Newton, the torque ξon a particle with radius\n1µm from optical trap is estimated to be ξ∼10−15N·m.\nConsidering that ˙ψis order of 10 GHz at most, the third\nterm in the bracket of Eq. (67) can be neglected.\nThe right hand side of Eq. (69) is same as the total\ntorque f(Ωz) defined in the main text, and vanishes in\nthe stable solutions of Ωz, which indicates\nΩz′≃˙ψ+ ˙φ= const . (72)\nSubstituting ˙ψ=Ωz′−˙φ, and treating Ωz′as constant\nvalue, Eq. (67)(68) are transformed to\nI1¨θ+β˙θ+/parenleftig\nI1˙φ˙ψ+ξ/parenrightig\nθ≃0, (73)\nI1θ¨φ+/parenleftbigg\n2I1˙θ+βθ−V\nγαM0θ/parenleftbigg\n1 +Ωz′\nω0/parenrightbigg/parenrightbigg\n˙φ\n≃/parenleftbigg\nI1˙θ−V\nγαM0θ/parenrightbigg\nΩz′, (74)\nup to the first order of ˙ φ. From Eq. (73), ˙ φcan be\nexpressed as\n˙φ=−1\nI1Ωz′θ/parenleftig\nI1¨θ+β˙θ+ξθ/parenrightig\n. (75)By substituting Eq. (75) into Eq. (74), we obtain\nI1...\nθ+ (2β+A)¨θ\n+/parenleftigg\nβ(β+A)\nI1+ 2ξ+I1Ω2\nz′+β˙θ\nθ+I1¨θ\nθ/parenrightigg\n˙θ\n+/parenleftbiggξ(β+A)\nI1+BΩ2\nz′/parenrightbigg\nθ= 0, (76)\nA=−V\nγαM0/parenleftbigg\n1 +Ωz′\nω0/parenrightbigg\n(>0), (77)\nB=−V\nγαM0(>0). (78)\nWe assume that the non-linear terms proportional to\n˙θ2/θand¨θ˙θ/θin Eq. (76) can be neglected (the validity\nof this assumption will be checked later). Then, Eq. (76)\nis reduced to a linear homogeneous differential equation\nof third order;\n...\nθ+a2¨θ+a1˙θ+a0θ= 0, (79)\na2= (2β+A)/I1(>0),\na1=/parenleftbiggβ(β+A)\nI1+ 2ξ+I1Ω2\nz′/parenrightbigg\n/I1(>0),\na0=/parenleftbiggξ(β+A)\nI1+BΩ2\nz′/parenrightbigg\n/I1(>0).\nAccording to Routh–Hurwitz stability criterion, the gen-\neral solution of Eq. (79) decays in time when and only\nwhen a2, a1, a0>0 and a2a1> a 0. Using I1≃9×\n10−27N·m·s2,β≃5×10−29N·m·s,ξ≃10−15N·m,\nB≃2.5×10−28N·m·s, and Ωz′∼1010Hz, we can\nconfirm a2a1> a0. Therefore, we can conclude that the\ndeclination angle θalways decays towards zero.\nFinally, we check the validity of the assumption of ig-\nnoring the non-linear terms in Eq. (76). Using the same\nparameter estimate shown above, Eq. (79) is approxi-\nmately rewritten as\n...\nθ+B\nI1/parenleftbigg\n1 +Ωz′\nω0/parenrightbigg\n¨θ+Ω2\nz′˙θ+B\nI1Ω2\nz′θ≃0. (80)\nThe general solution of this differential equation is writ-\nten as\nθ≃C1e−B\nI1t+ (C2sinΩz′t+C3cosΩz′t)e−BΩz′\n2I1ω0t,\n(81)\nwhere C1,C2, and C3are determined from the initial\nconditions for θ,˙θ,¨θ. If the initial condition is chosen to\nsatisfy ˙θ(t= 0)≪Ωz′θ(t= 0) and ¨θ≪Ω2\nz′θ(t= 0), C2\nandC3become much smaller than C1, since C1,C2Ωz′\nandC3Ω2\nz′are of the same order. Then, the amplitude of\nβ˙θ/θandI1¨θ/θbecome much smaller than a1≃I1Ω2\nz′,\nand neglecting these non-linear terms is justified.\nIn summary, we conclude that θdecays into zero and\nnever shows complex dynamics. This result indicates8\nthat the solution, Ω= (0,0, Ωz), which we derived in\nthe main article, is stabilized with respect to small per-\nturbation tilting the rotation axis.\n[1] Y. Arita, M. Mazilu, and K. Dholakia, “Laser-induced\nrotation and cooling of a trapped microgyroscope in vac-\nuum,” Nat. Commun. 4(2013).\n[2] Ren´ e Reimann, Michael Doderer, Erik Hebestreit,\nRozenn Diehl, Martin Frimmer, Dominik Windey, Fe-\nlix Tebbenjohanns, and Lukas Novotny, “Ghz rotation\nof an optically trapped nanoparticle in vacuum,” Phys.\nRev. Lett. 121, 033602 (2018).\n[3] Jonghoon Ahn, Zhujing Xu, Jaehoon Bang, Yu-Hao\nDeng, Thai M. Hoang, Qinkai Han, Ren-Min Ma, and\nTongcang Li, “Optically levitated nanodumbbell torsion\nbalance and ghz nanomechanical rotor,” Phys. Rev. Lett.\n121, 033603 (2018).[4] Fernando Monteiro, Sumita Ghosh, Elizabeth C. van As-\nsendelft, and David C. Moore, “Optical rotation of levi-\ntated spheres in high vacuum,” Phys. Rev. A 97, 051802\n(2018).\n[5] P. Huillery, T. Delord, L. Nicolas, M. Van Den Bossche,\nM. Perdriat, and G. H´ etet, “Spin mechanics with levitat-\ning ferromagnetic particles,” Phys. Rev. B 101, 134415\n(2020).\n[6] P. S. Epstein, “On the resistance experienced by spheres\nin their motion through gases,” Phys. Rev. 23, 710–733\n(1924).\n[7] A. Roth, Vacuum Technology (North Holland, 1990).\n[8] J. Corson, G. W. Mulholland, and M. R. Zachariah,\n“Calculating the rotational friction coefficient of fractal\naerosol particles in the transition regime using extended\nkirkwood-riseman theory,” Phys. Rev. E 96, 013110\n(2017).\n[9] This equation corresponds to Eq. (64) in Ref. 6.\n[10] This value for the molecular diameter ξis taken from\nFig. 2.9 of Ref. 7.\n[11] R. M. White, Quantum Theory of Magnetism (Springer,\nBerlin, Heidelberg, 2007)." }, { "title": "1609.07863v1.Ferromagnetic_resonance_study_of_composite_Co_Ni___FeCoB_free_layers_with_perpendicular_anisotropy.pdf", "content": "Ferromagnetic resonance study of composite Co/Ni - FeCoB free layers\nwith perpendicular anisotropy\nT. Devolder,1,a)E. Liu,2, 3J. Swerts,2S. Couet,2T. Lin,2S. Mertens,2A. Furnemont,2G. Kar,2and J. De\nBoeck2, 3\n1)Centre for Nanoscience and Nanotechnology, CNRS, Univ. Paris-Sud, Universit ´e Paris-Saclay, 91405 Orsay,\nFrance\n2)imec, Kapeldreef 75, 3001 Heverlee, Belgium\n3)Department of Electrical Engineering (ESAT), KU Leuven, Leuven 3001, Belgium\n(Dated: 18 June 2021)\nWe study the properties of composite free layers with perpendicular anisotropy. The free layers are made of a soft\nFeCoB layer ferromagnetically coupled by a variable spacer (Ta, W, Mo) to a very anisotropic [Co/Ni] multilayer\nembodied in a magnetic tunnel junction meant for spin torque memory applications. For this we use broadband ferro-\nmagnetic resonance to follow the field dependence of the acoustical and optical excitation of the composite free layer in\nboth in-plane and out-of-plane applied fields. The modeling provides the interlayer exchange coupling, the anisotropies\nand the damping factors. The popular Ta spacer are outperformed by W and even more by Mo, which combines the\nstrongest interlayer exchange coupling without sacrificing anisotropies, damping factors and transport properties.\nMagnetic tunnel junctions (MTJ) based on perpendicular\nmagnetic anisotropy (PMA) systems are under active devel-\nopment for the next generations of spin transfer torque (STT)\nmagnetic random access memories (MRAM). In this tech-\nnology, the information is stored in the magnetization state\nof a free layer composed typically of Fe-rich FeCoB alloys\nsandwiched between Ta and MgO (so-called ”single” MgO\nfree layer1) or between two MgO layers (”dual” MgO free\nlayers2). Each MgO interface provides an interface energy\nthat promotes PMA. The dual MgO option is gradually be-\ncoming more frequent for memory applications as it provides\nup to double anisotropy, hence more resilience to thermal\nfluctuations. This strategy ensures scalability3,4for junctions\ndown to 20-30 nm of diameter but material solutions have to\nbe found to scale further. To match with STT-MRAM ob-\njectives, one possible configuration comprises (i) an Fe-rich,\nbcc-structured, low damping FeCoB at the MgO interface (ii)\ncoupled ferromagnetically with a material supplying a strong\nanisotropy energy while maintaining damping and thickness\nas low as possible.\nAmong the material systems providing large PMA, the\nCo/Ni multilayers belong to the few ones that have a reason-\nably low damping, from50.033 down to6,70.021 and even8\n0.014 depending on compositions and thicknesses. The un-\nresolved questions are whether these low damping values can\nbe maintained in ultrathin Co/Ni multilayers and whether the\nfcc-based Co/Ni multilayer can be coupled strongly to the bcc-\nbased FeCoB layers across a texture transition.\nIn this paper, we compare three different spacers and\nevaluate their impact on the anisotropies, the interlayer\nexchange coupling and the damping in each layer of an\nMgO/FeCoB/spacer/[Co/Ni] \u00024/Pt system embodied in an\nMTJ. The studied spacers include the popular9Tantalum, as\nwell as Molybdenum and Tungsten spacers. The choice of bcc\nmetals (Ta, W and Mo) is to promote the fcc to bcc texture\na)Electronic mail: thibaut.devolder@u-psud.frtransition. The refractory character of W, Ta and Mo is also\na foreseen advantage since large resilience to atomic diffu-\nsion upon annealing is desirable for CMOS back-end-of-line\ncompatibility. W/CoFeB and Mo/CoFeB systems have indeed\nproven large resistance to annealing10,11which correlate with\nthe very slow diffusion of refractory metals in Fe12–14. To as-\nsess the performance of these spacers, we use broadband fer-\nromagnetic resonance (FMR) to follow the field dependence\nof the acoustical and the optical excitations of the compos-\nite free layer. The modeling provides the interlayer exchange\ncoupling, the anisotropies and the damping factors. The pop-\nular Ta spacer is outperformed by Mo, which combines the\nstrongest interlayer exchange coupling with no detrimental ef-\nfect on the anisotropies and the damping within the magnetic\nmaterials of the stack. This qualifies Molybdenum as a mate-\nrial of choice for the spacer layer in composite free layers.\nOur objective is to study hybrid free layers. By ”hybrid”\nwe mean comprising a bcc FeCoB layer which ensures op-\ntimal transport properties, coupled ferromagnetically to an\nfcc [Co/Ni] multilayer whose anisotropy strengthens thermal\nstability of the whole free layer. The multilayer is grown\nfirst on a Pt buffer that provides low coercivity and high\nanisotropy15. We then grow top-pinned MTJs of the follow-\ning configuration15: Pt / [Co(3 ˚A)/Ni(6 ˚A)]\u00024(t2= 3:5nm)\n/ spacer / Fe 60Co20B20(t1= 1 nm) / MgO / reference layer\n/ cap. The [Co/Ni] multilayer is terminated by a nickel layer\nin contact with the spacer. The studied spacers are Ta(3 ˚A),\nMo(3 ˚A) and W(3 ˚A). In addition, thicker (i.e. 5 ˚A) Mo and\nW spacers were used to decouple the two parts of the free\nlayer and thereby measure of their respective moments and\neasy axes. W(5 ˚A) leads to in-plane magnetization of the Fe-\nCoB layer (not shown). We aim to optimize the spacer within\nhybrid free layers in realistic MTJs, i.e. comprising reference\nlayers that might influence the crystallization within the free\nlayer. Our reference system15is a standard Co/Pt based syn-\nthetic ferrimagnet (SAF). All samples were annealed at 300\u000eC\nfor 30 min. in a field of 1 T.\nFor Mo(5 ˚A), the magnetizations are found to be Ms1=\n1:21\u0002106A=m,Ms2= 0:763\u0002106A=mfor the FeCoBarXiv:1609.07863v1 [cond-mat.mtrl-sci] 26 Sep 20162\n/s48/s50/s48/s52/s48/s54/s48\n/s48/s50/s48/s52/s48/s54/s48\n/s45/s50 /s45/s49 /s48 /s49/s48/s50/s48/s52/s48/s54/s48\n/s45/s49 /s48 /s49/s45/s50 /s45/s49 /s48 /s49 /s50/s32/s32/s69/s105/s103/s101/s110/s109/s111/s100/s101/s32/s102/s114/s101/s113/s117/s101/s110/s99/s105/s101/s115/s32/s40/s71/s72/s122/s41/s78/s111/s32/s99/s111/s117/s112/s108/s105/s110/s103/s69/s102/s102/s101/s99/s116/s105/s118/s101/s32/s97/s110/s105/s115/s111/s116/s114/s111/s112/s105/s101/s115\n/s70/s101/s67/s111/s66/s58/s32 /s51/s50/s49/s32/s38/s32 /s67/s111/s47/s78/s105/s58/s49/s32/s50/s56/s55/s32/s109/s84\n/s32/s32\n/s74/s61/s48/s46/s57/s32/s109/s74/s47/s109/s50/s32/s32\n/s74/s61/s48/s46/s53/s32/s109/s74/s47/s109/s50/s45/s54/s32/s38/s32 /s49/s48/s49/s56/s32/s109/s84\n/s32/s32/s32/s32\n/s79/s117/s116/s32/s111/s102/s32/s112/s108/s97/s110/s101/s32/s102/s105/s101/s108/s100/s32/s40/s84/s41\n/s69/s97/s115/s121 /s32/s97/s120/s105/s115/s32/s73/s110/s32/s112/s108/s97/s110/s101/s32/s102/s105/s101/s108/s100/s32/s40/s84/s41\n/s72/s97/s114/s100/s32/s97/s120/s105/s115/s45/s50/s51/s32/s38/s32 /s52/s52/s48/s32/s109/s84/s32/s76/s111/s111/s112/s115 /s69/s105/s103/s101/s110/s109/s111/s100/s101/s115\nFIG. 1. (Color online). Eigenmode frequencies and hard axis loops for 3 hybrid free layers that share the same calculated eigenexcitation\nfrequencies at remanence. The anisotropies and the interlayer couplings are chosen to get the eigenmode frequencies of 9 GHz (blue dot)\nand 36 GHz (red dot) at remanence like in the case of Mo(3 ˚A) spacer. Left column: eigenmode frequencies versus out-of-plane fields for an\nincreasing (black) or decreasing (green) field sweep. In practice the layers’ magnetizations switch near zero field and thus visit mainly the\ncolored branches that correspond to the most stable states. Middle column: idem versus in-plane fields. Right column: normalized hard axis\nloops (i.e.MxversusHx) of the FeCoB layer (black lines) and the [Co/Ni] multilayer (red lines).\nlayer and the [Co/Ni] multilayer, respectively. For the thinner\nspacers, the two parts of the free layer are sufficiently coupled\nto switch as a single block during easy axis loops, while the\nSAF reference system switches in two steps at much larger\npositive and negative fields [Fig. 2(a)]. Hard axis loops (not\nshown) comprise also the signals of the SAF of the MTJ such\nthat the informations are too intertwined for a separate identi-\nfication of the contribution of each magnetic sub-block.\nTo see to what extent the two parts of the free layer ful-\nfill their roles, we measured the system’s eigenexcitations, in\naddition its ground state determined by magnetometry. We\nillustrate our method in Fig. 1. We are concerned by hybrid\nfree layers with PMA and full remanence, i.e. in which the\nmagnetizations of the two layers at remanence are collinear to\nthe out-of-plane axis (z). In this case the acoustical and opti-\ncal eigenmode frequencies of the free layer under out of plane\napplied fields17arefexp\n1;2=\r0\n2\u0019(Hsoft\n1;2\u0006Hz)where the fields\nadd for magnetizations up [i.e. along +(z)] and subtract for\nmagnetizations down. The softening fields read17:\nHsoft\n1;2=He\u000b\nk1+He\u000b\nk2\n2+J\n2Ms2t2+J\n2Ms1t1\u0006p\n\u0001\n2Ms1Ms2t1t2(1)\nwhere \u0001 = 2 JMs1Ms2t1t2(He\u000b\nk2\u0000He\u000b\nk1)(Ms1t1\u0000\nMs2t2) +Ms12Ms22t12t22(He\u000b\nk2\u0000He\u000b\nk1)2+J2(Ms1t1+Ms2t2)2. We have used the notation He\u000b\nk1;2for the effective\nanisotropy fields of FeCoB (layer 1) and Co/Ni (layer 2), and\nJfor their interlayer exchange coupling.\nWhen later conducting FMR versus Hzexperiments on the\nfree layer (Fig. 2), we will confirm the finding of two V-shaped\nlines, with the apex of the V at fexp\n1;2and slopes\u0006\r0\n2\u0019. Un-\nfortunately, the two data ( fexp\n1andfexp\n2) are insuffisient to\ndetermine the three unknown parameters fJ; He\u000b\nk1; He\u000b\nk2g. We\nhave illustrated this problem in Fig. 1 (left panels) by choos-\ning three different triplets fJ; He\u000b\nk1; He\u000b\nk2gthat yield the same\ncalculated eigenexcitation frequencies in the Hzfield config-\nuration (see the red and blue lines common to all left panels).\nTo get an overdetermined problem, we performed in addi-\ntion FMR vs in-plane field. Indeed as the two parts of the free\nlayer have distinct anisotropies and thickness, their magneti-\nzations tilt at different angles and the dynamics gets very sen-\nsitive to the exchange coupling J. The fitting procedure can\nbe summarized this way: we first apply Eq. 1 on the two re-\nmanent frequencies fexp\n1;2to get all the triplets fJ; He\u000b\nk1; He\u000b\nk2g\ncompatible with the easy axis field data. To determine which\ntriplet is the correct one, we numerically calculate the mag-\nnetic configuration and eigenexcitations in in-plane applied\nfield for all possible triplets, (Fig. 1, right panels) and select\nthe triplet that best match with the experiments.3\nTABLE I. Properties of the free layer subsystems. The magnetizations were fixed at Ms1= 1:21\u0002106A=m,Ms2= 0:763\u0002106A=m. The\nysymbol emphasizes the number that come from high field extrapolation when the dispersion curve departs from linear behavior at low fields.\nThe linewidths are calculated from out-of-plane measurements of the lowest and highest frequency eigenmodes. In the absence of coupling,\nthey would be indicative of the damping of the FeCoB and Co/Ni systems, respectively.\nfexp\n1fexp\n2 Fe60Co20B20[Co3 ˚A/Ni6 ˚A]\u00024Coupling1\n2@\u0001f\n@f1\n2@\u0001f\n@fTMR RA\n(GHz) (GHz)\u00160He\u000b\nk1\u00160He\u000b\nk2J(mJ/m2)FeCoB Co/Ni\nSpacer \u000630 mT\u000650 mT\u00060.01\u00060:001 \n:\u0016m2\nTa 3 ˚A 1.0 24.5 -20 850 0.07 0.009 0.03\u00060:004 137% 7.0\nMo 3 ˚A8.6 35 -50 920 0.58 0.010 0.03\u00060:006 137% 8.7\nW 3 ˚A-0.7y27.6 -190 890 0.22 0.014 0.030\u00060:003 137% 8.7\nxxxxRef. \nRef. Ref. layersFree. layer\nFIG. 2. (Color online). Properties of the MTJ with Mo 3 ˚A spacer. (a)\nEasy axis loop of the MTJ. (b) Permeability in the ffrequency-fieldg\nplanes for (b) out-of-plane field and (c) in-plane applied fields. The\ncontrast has been strengthened in the green box. The violet arrows\npoint at an eigenmode of the MTJ’s reference layer. (d) Eigenmode\nfrequencies along a hard axis loops (symbols) and modeling thereof\n(lines) with the parameters of Table 1.\nIn the absence of coupling [Fig. 1, top middle panel], the\nhard-axis field eigenmode lines are W-shaped with two dis-\ntinct softening fields at Hx=\u0006He\u000b\nk1for the FeCoB layer and\natHx=\u0006He\u000b\nk2for the Co/Ni multilayer. For finite coupling\nJ, an anticrossing appears in the dispersion curves and there\nremains only one softening field per applied field sign. The\nlayers still tilt at a different pace with the field. For large cou-\npling (bottom panels, J= 0:9 mJ=m2), the layers essentially\ntilt together and behave like a single unit.\nIn practice the shape of the in-plane field FMR curves is a\nway to select the set of material parameters that best describe\na sample. Besides, the linewidth of each mode in the out-of-plane field configuration can be used to estimate the damping\nof the two parts of the free layer. Indeed we will see that the\ncoupling is small enough so that the lowest mode linewidth re-\nflects the value of the FeCoB damping, while the linewidth of\nthe highest frequency mode reflects the damping of the Co/Ni\npart of the free layer (for a detailed justification see for in-\nstance Figs. 6 and 7 in ref. 17 and the related analysis).\nThe free layer eigenmodes were determined experimen-\ntally using Vector Network Analyzer FerroMagnetic Reso-\nnance (VNA-FMR18). We applied fields either in the plane of\nthe sample or perpendicularly to it. Thanks to the very differ-\nent anisotropies of the magnetic subsystems within the MTJ,\nthe eigenmodes have well separated frequencies [Fig. 2(b) and\n(c)]. There is one eigenmode undergoing sudden frequency\njumps at exactly the two switching fields of the SAF reference\nlayers of the MTJ, but not undergoing any change at the free\nlayer coercivity. The frequency of this mode is independent\nof the free layer inner spacer (i.e. Mo, W or Ta). This mode\ncan thus be assigned to the other part of the MTJ (i.e. to the\nSAF reference system), and we shall thus not consider here-\nafter. Following the methods described earlier (Fig. 1), a fit of\nthe other modes to coupled macrospins was used to determine\neach layer’s properties with conclusions gathered in Table I.\nfexp\n1describes the strength of the effective fields that hold\nthe softest part of the free layer. It is thus a relevant indica-\ntion of its non volatility. The hybrid free layer with Mo spacer\nclearly outperforms substantially the other spacers in term of\nnon volatility. Mo also ensures a low damping of the FeCoB\nlayer at no expense of the overall anisotropy and of the trans-\nport properties. The main influence of the chemical nature of\nthe spacer is the coupling strength (Table 1). The magneto-\ntransport properties are insensitive to the spacer layer. The\ncouplings through the Ta spacer is low, and our experience\nis that it is weaker in our present Ni-terminated multilayers\nthan when Co-terminated19,20. Comparatively to Ta, W could\nbe considered as better owing to its larger interlayer exchange\ncoupling. However, the W spacer has a clear detrimental ef-\nfect on both the damping and the anisotropy of the FeCoB\nlayer, which is consistent with other studies21.\nIn summary we have studied hybrid perpendicular\nanisotropy free layers that couple a soft FeCoB layer with a\nvery anisotropic [Co/Ni] multilayer through various spacers\nmade of refractory metals. The formerly used Ta(3 ˚A) spacer\nis outperformed by W(3 ˚A) and even more by Mo(3 ˚A) spac-\ners, which combine the strongest interlayer exchange coupling\n(0.58 mJ/m2) without sacrificing the anisotropies, the damp-4\ning factors and the magneto-transport properties within the\nstack.\n1S. Ikeda, K. Miura, H. Yamamoto, K. Mizunuma, H. D. Gan, M. Endo,\nS. Kanai, J. Hayakawa, F. Matsukura, and H. Ohno, Nature Materials 9,\n721 (2010).\n2H. Sato, M. Yamanouchi, S. Ikeda, S. Fukami, F. Matsukura, and H. Ohno,\nApplied Physics Letters 101, 022414 (2012).\n3H. Sato, E. C. I. Enobio, M. Yamanouchi, S. Ikeda, S. Fukami, S. Kanai,\nF. Matsukura, and H. Ohno, Applied Physics Letters 105, 062403 (2014).\n4J.-H. Kim, J.-B. Lee, G.-G. An, S.-M. Yang, W.-S. Chung, H.-S. Park, and\nJ.-P. Hong, Scientific Reports 5, 16903 (2015).\n5W. Chen, J.-M. L. Beaujour, G. d. Loubens, A. D. Kent, and J. Z. Sun,\nApplied Physics Letters 92, 012507 (2008).\n6H.-S. Song, K.-D. Lee, J.-W. Sohn, S.-H. Yang, S. S. P. Parkin, C.-Y . You,\nand S.-C. Shin, Applied Physics Letters 103, 022406 (2013).\n7H.-S. Song, K.-D. Lee, J.-W. Sohn, S.-H. Yang, S. S. P. Parkin, C.-Y . You,\nand S.-C. Shin, Applied Physics Letters 102, 102401 (2013).\n8M. Haertinger, C. H. Back, S.-H. Yang, S. S. P. Parkin, and G. Woltersdorf,\nJournal of Physics D: Applied Physics 46, 175001 (2013).\n9L. Cuchet, B. Rodmacq, S. Auffret, R. C. Sousa, and B. Dieny, Applied\nPhysics Letters 105, 052408 (2014).\n10T. Liu, Y . Zhang, J. W. Cai, and H. Y . Pan, Scientific Reports 4, 5895\n(2014).11Y . Liu, T. Yu, Z. Zhu, H. Zhong, K. M. Khamis, and K. Zhu, Journal of\nMagnetism and Magnetic Materials 410, 123 (2016).\n12R. W. Powers and M. V . Doyle, Journal of Applied Physics 30, 514 (1959).\n13S. Takemoto, H. Nitta, Y . Iijima, and Y . Yamazaki, Philosophical Magazine\n87, 1619 (2007).\n14H. Nitta, T. Yamamoto, R. Kanno, K. Takasawa, T. Iida, Y . Yamazaki,\nS. Ogu, and Y . Iijima, Acta Materialia 50, 4117 (2002).\n15E. Liu, J. Swerts, S. Couet, S. Mertens, Y . Tomczak, T. Lin, V . Spamp-\ninato, A. Franquet, S. V . Elshocht, G. Kar, A. Furnemont, and J. D. Boeck,\nApplied Physics Letters 108, 132405 (2016).\n16M. Gajek, J. J. Nowak, J. Z. Sun, P. L. Trouilloud, E. J. OSullivan, D. W.\nAbraham, M. C. Gaidis, G. Hu, S. Brown, Y . Zhu, R. P. Robertazzi, W. J.\nGallagher, and D. C. Worledge, Applied Physics Letters 100, 132408\n(2012).\n17T. Devolder, Journal of Applied Physics 119, 153905 (2016).\n18C. Bilzer, T. Devolder, P. Crozat, C. Chappert, S. Cardoso, and P. P. Freitas,\nJournal of Applied Physics 101, 074505 (2007).\n19T. Devolder, S. Couet, J. Swerts, and A. Furnemont, Applied Physics Let-\nters108, 172409 (2016).\n20A. Le Goff, R. Soucaille, T. Tahmasebi, J. Swerts, A. Furnemont, and\nT. Devolder, Japanese Journal of Applied Physics 54, 090302 (2015).\n21R. Soucaille, M. Belmeguenai, J. Torrejon, J.-V . Kim, T. Devolder, Y . Rous-\nsign, S.-M. Chrif, A. A. Stashkevich, M. Hayashi, and J.-P. Adam,\narXiv:1604.05475 [cond-mat] (2016), arXiv: 1604.05475." }, { "title": "1501.02938v1.Strain_Induced_Extrinsic_High_Temperature_Ferromagnetism_in_the_Fe_Doped_Hexagonal_Barium_Titanate.pdf", "content": " \n1 \n Correspondence and requests for materials should be addressed to A.Z. (andrej.zorko@ijs.si) \n \nStrain -Induced Extrinsic High-\nTemperature Ferromagnetism in the Fe -\nDoped Hexagonal Barium Titanate \nA. Zorko1,*, M. Pregelj1, M. Gomilšek1, Z. Jagličić2,3, D. Pajić4, M. Telling5,6, I. Arčon1,7, I. \nMikulska7, and M. Valant7 \n1Jožef Stefan Institute, Jamova c. 39, SI -1000 Ljubljana, Slovenia , 2Institute of Mathematics, \nPhysics and Mechanics, Jadranska c. 19, 1000 Ljubljana, Slovenia , 3Faculty of Civil and Geodetic \nEngineering, University of Ljubljana, Jamova c. 2, SI -1000 Ljubljana, Slovenia , 4Department of \nPhysics, Faculty of Science, University of Zagreb, Bijenička c. 32, HR-10000 Zagreb, Cr oatia, \n5ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 OQX, United \nKingdom , 6Department of Materials, University of Oxford, Parks Road, Oxon, UK , 7University of \nNova Gorica, Vipavska 13, SI -5000 Nova Gorica, Slovenia . \n \nDiluted magnetic semiconductors possessing intrinsic static magnetism at high \ntemperatures represent a promising class of multifunctional materials with high \napplication potential in spintronics and magneto -optics . In the hexagonal Fe -doped \ndiluted magneti c oxide , 6H-BaTiO 3-, room -temperature ferromagnetism has been \npreviously reported . Ferromagnetism is broadly accepted as an intrinsic property of \nthis material , despite its unusual dependence on doping concentration and \nprocessing conditions . However, t he here reported combination of bulk \nmagnetization and complementary in-depth local -probe electron spin resonance and \nmuon spin relaxation measurements , challenges this conjecture . While a \nferromagnetic transition occurs around 700 K, it does so only in ad ditionally \nannealed samples and is accompanied by an extremely small average value of the \nordered magnetic moment . Furthermore, several additional magnetic instabilities \nare detected at lower temperatures. These coincide with electronic instabilities of \nthe Fe-doped 3C-BaTiO 3- pseudocubic p olymorph . Moreover , the distribution of iron \ndopants with frozen magnetic moments is found to be non-uniform. Our results \ndemonstrate that the intricate static magnetism of the hexagonal phase is not \n2 \n intrinsic, but rather stems from sparse strain -induced pseudocubic regions . We \npoint out the vital role of internal strain in establishing defect ferromagnetism in \nsystems with competing structural phases. \nThe search for dilute magnetic oxides (DMOs) is at the forefront of spintronic s and magneto -\noptics research and application1–6. The great interest has been triggered by the possibility of \ncombining diverse functionalities of semiconductor electronic s and magneti sm in a single \nmaterial , which would boost its application potential . Existence of magn etoelectric coupling in \nwide-band-gap semiconducting oxides at room temperature would be of great technological \nimportance , given that the magnetic ordering is indisputably intrinsic. Indeed, a number of \nstudies have claimed such intrinsic ordering has been observed in non-magnetic oxide \nsemiconductors lightly doped with paramagnetic transition -metal ions. These findings are, \nhowever, a subject of controversy7–9. Very often the corresponding studies are focused only \non routine characterization and modelling of the material's properties rather than critical \nanalysis of the origin of the observed static magnetism . Recently, more detailed experiments \nhave proven wrong a significant part of the claims about intrinsic magnetism in D MOs10–14. \nMoreover, a ggregation of magnetic ions that leads to chemical phase separation on a nano -\nscale, also known as spinodal decomposition into regions with high/low concentration of \ndopants and a crystal structure imposed by the semiconducting host, has been lately \nwitnessed in various cases2,3,15-17. Thus, the need for careful analysi s of the magnetic \nproperties of these systems at the microscopic level and their critical assessment are of \nparamount importance . \n \nSince the discovery18 of room -temperature ferromagnetism in the bulk Fe-doped hexagonal19 \n6H-BaTiO 3- (Figure 1) that can be stabilized from the pseudocubic (3C) perovskite struct ure \n(Figure 1) by transition -metal doping20, this system has been in the focus of investigation s. \nAlthough simultaneous mag netic and polar orders (multiferroicity), highly desired for device \napplications , cannot be achieved in this ma terial (1% Fe doping destroys ferroelectric \norder18), the magnetoelectric coupling has recently been demonstrated21. Moreover, \narguments speaking in favour of its intrinsic ferromagnetism would make Fe -doped 6H- \n3 \n BaTiO 3- rather special in the field of ferromagnetic (FM) DMOs. In contrast to many other \nmaterials where dopant -ion segregation and formation of precipitates a re regularly \nencountered, the very high solubility of 6 H-BaTiO 3 for Fe (and other transition metals ), the \npossibility of growing single -crystal DMOs, and the apparent absence of any spurious \nprecipitates all seem as strong evidences for its intrinsic FM character18,22,23. \n \nFigure 1 : Schematic interface of the two crystallographic phases in BaTiO 3-. \nPseudocubic (3 C) and hexagonal (6 H) crystallographic polymorph of BaTiO 3-Fe3+ cations \nsubstitute for Ti4+ ions (spheres) within FeO 6 ochatedra (polyhe dra). Two crystallographically \ndifferent Ti sites are found in the 6 H structure, while only one exists in the 3 C structure. High-\ntemperature ferromagnetism (FM) is ascribed to sparse regions of the 3 C phase, while the \nmajority 6 H phase remains paramagnetic (PM). \nThe FM state of the Fe -doped 6 H-BaTiO 3- is, however, highly unusual, if not even \ncontroversial . While t he amount of the room -temperature FM signal significantly decreases \nwith increasing dopant concentration24, the average value of the static magnetic moment per \nFe3+ ion never exceeds a few hundredth s of a Bohr magneton ( B); compared to the full \nmoment of 5 B. Different hypotheses about such a drastic moment reduction have been put \nforward in the literature . These include competition between s uperexchange interactions \nleading to intrinsic magnetic inhomogeneity in the system18 or, alternatively, the formation of \ndispersed Fe3+ clusters around oxygen vacancies24. Furthermore, the size of the average FM \nmoment substantially depend s on various factors, including temperature, atmosphere21,25, and \nduration of the synthesis26. Also the inevitable oxygen vacancies in the Fe-doped 6 H-BaTiO 3- \n \n4 \n material, which compensate for the charge mismatch between Fe3+ and Ti4+, are \nsuggested21,23 to play an intrinsic role in establishing ferromagnetism through a dynamical \nexchange of trapped e lectrons among bound polarons27. However, the corresponding reports \nare contradictory , since the FM response can apparently be enhanced in either reduced \nsamples22 or in samples annealed in oxygen atmosphere25. Moreover, it has recently been \nsuggested that the FM behaviour of the Fe -doped 6 H-BaTiO 3- should rather be related to \nthermally activated di ffusion , leading to a cation order -disorder transition , and that the \nsegregation of oxygen vacancies plays no major role in ferromagnetism28. The result ing \nclustering of Fe3+ ions at the two neighbouring pentahedral Ti(2) sites (Figure 1) and the \nsegregation of oxygen vacancies at the connecting oxygen site29, lead to the formation of \nlattice -matched Fe 2O8 dimers . These have been predicted by ab -initio calculations18, and are \nregularly observed in experiment s28,30,31, further revealing that a major portion of the Fe \ndopants are involved in their formation28,31. This fact thus evidently questions the cation \nordering as the intrinsic origin of ferromagnetism in 6 H-BaTiO 3-, given the extremely small \nsize of the average value of the magnetic moment per Fe dopant. \n \nThe FM behaviour of the Fe-doped 6H-BaTiO 3- has not yet been explored above room \ntemperature or examined by local -probe magnetic spectroscopy techniques , which could \npotentially help in identifying any spurious effects. Therefore, we performed a comprehensive \nmagnetic investigation, combin ing both bulk magnetic measurements as well as the \ncomplementary local -probe measurements of electron spin resonance (ESR) and muon spin \nrelaxation ( SR), in the broad temperature range between 2 and 850 K. Here, we report a \nplethora of magnetic anomalies that appear in the Fe -doped 6 H-BaTiO 3- below ~700 K. These \ncoincide with the electronic instabilities of the Fe -doped pseudocubic 3 C-BaTiO 3 polymorph. \nWe provide firm experimental evidence that the previously reported static magnetism is not \nan intrinsic property of the doped 6 H-BaTiO 3 hexagonal phase, but is instead related to sparse \nregions of strain -induced pseudocubic phase , which arise as a consequence of locally \ncompeting structural phases . \nResults \n5 \n In order to pr ovide clear evidence for either intrinsic or extrinsic origin of magnetism in the \nFe-doped 6 H-BaTiO 3-, an in-depth experimental study , which combin es bulk and local -probe \nmagnetic characterization techniques, is necessary. All our measurements were performed on \nhigh-quality polycrystalline samples (see Methods ); some of which were additionally \nannealed . This annealing prove s crucial for the existence of FM behaviour . In what follows, we \nlabel the non -annealed and annealed samples with F cBTO and F cBTOa, respectively . Here c \ndenotes the doping percent age of Fe3+ ions at the Ti sites. \n \nDete ction of magnetic instabilities by b ulk magnetic measurements \nBulk magnetization (M) measurements display a clear difference between the non-annealed \nand annealed samples (Figure 2a). While t he latter exhibit complicated zero -field cooled \n(ZFC)/field -cooled (FC) split M(T) curves , a paramagnetic (PM) phase is found in the former . \nThe PM phase is characterized by a monotonic temperature dependence and absence of the \nZFC/FC bifurcation . The corresponding molar susceptibility (per mole Fe) for the non -annealed \nsamples, \nmol MH (H denotes the applied magnetic field ), agreeswell with the Curie -Weiss \nmodel \n 2\nmol A B CW 3 N k T . Here NA and kB denote the Avogadro and the Boltzman \nconstant, respectively, CW is the Weiss temperat ure and = 4.6(1) B is the average magnetic \nmoment per Fe dopant. The derived magnetic moment is common to both F10BTO and \nF20BTO samples and is somewhat reduced from th e value of \n BB 1 5.92 g S S expected \nfor Fe3+ (S = 5/2) moments. However, the Fe3+ valence was clearly revealed by X -ray \nabsorption near -edge structure (XANES) spectroscopy28,29 as being by far dominant in 6 H-\nBaTiO 3-. For F10BTO and F20BTO samples we find antiferromagnetic (AFM) Weiss \ntemperature of -3(1) K and -7(2) K, respectively. This parameter scales linearly with the \ndoping concentration and provides the energy scale of the average excha nge coupling J \nbetween the iron dopants . We find J to be rather small compared to room temperature , where \nthe FM state has been reported . Secondly, it is AFM and thus questions the intrinsic origin of \nthe FM order. \n \n6 \n At high temperatures (T > Tc1 ~ 700 K) the M(T) curves of the anneal ed samples match those \nof the non -annealed ones, proving the same PM state. This behaviour , however, drastically \nchanges below Tc1, especially in the low applied field of 10 mT (Figure 2a). On decreasing \ntemperature a ZFC/FC splitting is observed in F20BTOa at Tc1, with the FM component being \nenhanced below Tc2 ~ 570 K, and the ZFC/FC di fference increased below Tc3 ~ 450 K . Finally, \nan additional magnetic anomaly , reflected in a local magnetization maximum , is found at Tc4 \n~ 110 K. The M(H) magnetization curves of the same sample (Figure 2b) clearly display the \nsame high-temperature magnetic transitions , as a hysteresi s appears below Tc1 and is \nenhanced below Tc2. The deduced average value of the FM ordered magnetic moment (FM = \n1.5∙10-2B per Fe at 300 K; see Methods ) is dramatically reduced from the expected full 5B \nFe3+ value . Moreover, the M(H) curves change linearly with H at larger fields; a response \nbeing typical for PM and not FM systems . Therefore, below Tc1 one can think of the system as \nbeing predominantly PM with an additional small FM component. Such a simple picture is \nconfirmed by M(T) data in higher magnetic fields, where the high -temperature magnetic \nanomalies, although appearing at the same critical temperatures, become drastically \nsuppressed (Figure 2a) due to saturation of the FM component . Similarly, as regularly \nreported before18,23, the PM component that shows a Curie -Weiss temperature dependence \ndominates the magnetic response at low -temperature s (see the inset in Figure 2b), as the \nsplitting of the FM hysteresis is only moderately enhanced at 2 K compared to room \ntemperature (Figure 2b). The PM component is far from being saturated at low temperatures \neven at 5 T , similarly as found before18,22,23. The average value of the PM magnetic moment \nper Fe3+ ion can be modelled32 with a modified Brillouin function for spin 5/2 , \n5/2 B effB M f B k T \n, by renormalizing the moment (\nB 52g ; g denotes the Landé g-\nfactor) with a factor f and introducing an effective temperature \neff 0T T T . The fit to the \nFBTO20a data yields the empirical temperature T0 = 2.1 K (inset in Figure 2b) , confirming \nthat AFM interactions are pertinent to this phase32. As the actual concentration of iron dopants \nin all our samples is very close to the nominal one (see Methods), t he obtained \nrenormalization factor f = 0.25 reveals that the majority of spins do not contribute to the \nmagnetic signal at low temperatures. Such renor malization factor that has been found to \n7 \n decrease steadily with the doping concentration18 can then be accounted for by formation of \nnon-magnetic AFM entangled iron pairs at low temperatures , which are correlated with the \nformation of the Fe2O8 structural dimers28,31. The bulk-magnetization results thus imply the \npresence of FM and PM magnetic phases in the annealed 6H -BaTiO3 - samples below Tc1 and \ncorroborate impurity -based ferromagnetism, as further verified below by local -probe magnetic \ncharacterization s. \n \nFigure 2: Bulk magnetization measurements. (a) The temperature dependence of zero -\nfield cooled (full symbols) and field -cooled (open symbols) field -normalized bulk magnetization \nin the 20% Fe -doped 6 H-BaTiO 3-. The solid line is the fit to the Curie -Weiss model \n2\n0 A B CW 3 M H N k T \n (see text for details ), while the vertical dashed lines display the \ntemperatures of the magnetic transitions in the annealed sample. (b) Magnetization curves of \nF20BTOa at different temperatures displaying the FM hysteresis . Inset shows the 2 K data \n(circles) and the fit with the modified Brillouin function \n5/2 B effB M f B k T (see text for \ndetails) . (c) The time -dependent magnetization in the F20BTOa sample at 300 K after applying \nand rem oving the magnetic field of 1 T . \nPeculiarities of FM resonance modes \nFurther insight to the unusual magnetism of the Fe -doped BaTiO 3- samples on a local scale is \nprovided by ESR. Again, dra stically di fferent ESR spectra are found for the non -annealed and \nannealed samples . The former are featureless single -line ESR spectra , while the latter possess \nseveral lines that significantly evolve with temperature below Tc3 (Figure 3a). We note that \n0 200 400 600 8000.00.10.20.30.40.50.6\n-0.10 -0.05 0.00 0.05 0.10-0.4-0.20.00.20.4\n-1000100-303\n0 10 20 30 400.010.111\nTc1Tc2Tc3Tc4\n M/0H (Am2 T -1 per mol Fe)\nT (K)a\n F20BTO\n 10 mT F20BTOa\n 10 mT\n 0.34 T,\n,F20BTOa\n2 K400 K\n850 K800 K700 K600 K500 Kb\n M (10-2B per Fe)\nBH (T)data\n fit \n \n \n2 KF20BTOac\n0 T0 T1 T M (10-2B per Fe)\nt (min) \n8 \n the ESR spectra of the PM non-annealed samples are different from the fine-structured ESR \nspect ra reported in previous studies22,33 –35. However, it should be stressed that all previous \nstudies were limited to doping concentrations below 5%. The disappearance of the fine \nstructure in our non -annealed samples suggests enhanced magnetic interactions (dipolar \nand/or exchange) between iron dopants . \n \nFigure 3: ESR characterization of 6 H-BaTiO 3-. (a) The temperature dependence of the ESR \nspectra in 20% Fe -doped 6 H-BaTiO 3- displaying a single line in the non -annealed sample and \na multi -feature line in the annealed sample . Spectra are displaced vertically for clarity. The g = \n2 position is marked with the vertical line. (b) The increase of the ESR line width with doping \nconcentration in the non -annealed samples at 300 K . The peak -to-peak line width B of both \nspectr a is indicated by the horizontal arrow s. (c) Comparison of the 300 -K spectra of F20BTOa \nat 9.4 GHz and at 24.3 GHz . Side -band shifts that are frequency independent are marked with \nvertical lines. ( d) Resonance field of various resonance modes observed in the annealed Fe -\ndoped 6 H-BaTiO 3- samples below Tc3. The temperature independent position (horizontal mode) \ncorresponds to the g = 2 mode. (e) Comparison of the ESR intensity decrease of the single -line \n0 200 400 600-6061218243036\n-200 0 200 400-0.50.00.5\n0 200 400 600-0.9-0.6-0.30.00.30.6\n0 200 400 6000.00.20.40.6\n450 500 550 6000.00.20.40.60.81.0g = 2600 K\n500 K (x 1/2)\n400 K\n300 K\n200 K\n100 K\n5 K (x 1/2)\nF20BTOa\nF20BTOa\n ESR signal (arb. units)\n0H (mT)c\n0(H - H0) (mT) F20BTOa\n9.4 GHz\n24.3 GHzF10BTO\nF20BTO\n0H (mT) \n b\nB\ng = 2e\nT (K)T 20%\nc3T 10%\nc3Tc4d\n Bres (T)F10BTOa\nF20BTOa\n F20BTOa\n F20BTOESR int. (arb. units)\nT (K) \nTc3Tc2 \n9 \n FM mode observed above Tc3 in F20BTOa and the PM signal of the F20BTO sample with Curie -\nlike dependence. Both ESR intensities are normalized at 450 K. \n \nIn order to support the claim of sizable magnetic interactions we analyse the line width of the \nPM ESR spectra in the non -annealed samples. The peak -to-peak line width B (Figure 3b) \nscales proportionally with the doping level (\n10%45 2 mT B ,\n20%91 2 mT B ). Such a \nlinear increase is in accordance with the linearly increasing magnetic anisotropy and exchange \ninteractions (see Methods), as deduced from the Weiss temperatures of both non -annealed \nsamples. On the contrary, the dipolar interactions in the absence of isotropic exchange \ncouplin g36 would yield much smaller line widths (see Methods). ESR thus verifies that the \naverage exchange interaction J among dopants is not negligible. Furthermore, the line width, \nbeing proportional to c, is a fingerprint of a uniform PM phase. Aggregation of magnetic ions \ninto nano -scale clusters, on the other hand, would result in an inverse dependence of the ESR \nline width on c, as was recently observed17 in Mn -doped SrTiO 3. \n \nIn sharp contrast to the simple ESR line shape of the PM non -annealed samples, the ESR \nspectra of the annealed samples are much more complex below Tc3. Here, they exhibit several \nwell-displaced lines (Figure 3c) that shift significantly with temperature (F igure 3d). The field \nshift of the side bands from the g = 2 position is the same at 9.4 GHz and at 24.3 GHz (Figure \n3c), which is indicative of static internal magnetic fields characteristic of the FM phase. Very \nimportantly, above Tc3 a single -line spectrum is observed at \n2g (Figure 3a) in all the \nannealed samples despite the fact that the FM state persists to much higher temperature, Tc1 \n~ 700 K. With its peak-to-peak line width \n9 1 mTB , this spectrum is markedly narrower \nthan the PM spectrum of the non -annealed samples . Moreover, its intensity strongly \ndecreases between Tc3 and Tc2 (Figure 3e). The latter behaviour is in disagreement with the \nCurie-like decrease, characteristic of the PM non -annealed samples (Figure 3e) , and is rather \nreminiscent of the FM magnetization decrease in the same temperature range (Figure 2a). \nTherefore, this \n2g mode is still a FM mode. We stress that its single -component line shape \n10 \n reveals that the local magnetic anisotropy, which dictates the shape of the ESR spectrum in \nthe FM state37 and depends on local symmetry, drastically ch anges at Tc3. \n \nClustering of FM moments as evidenced by SR \nThe ESR measurements provide convincing evidence of FM moments in the annealed samples , \nas well as their absence in the non -annealed samples . In addition , they reveal a novel and \nimportant (vide infra) discovery about the local -symmetry reduction below Tc3. However, ESR \ndoes not yield any information about the spatial distribution of the FM moments. T o answer \nthis question we turn to the complementary SR method (see Methods for introduction to th is \nmethod ). This local-probe technique is renown ed in magnetic spectroscopy for its ability to \ndetect even the smallest local magnetic fields at the muon stopping site and for the ease in \nwhich the experimentalist can discriminate between static and dynamical internal fields38. We \nfind that i n zero magnetic field (ZF) , the time -dependent muon spin polarization P(t) is \nmonotonic at all temperatures in the experimental time window, independe nt of the PM/FM \ncharacter of a particular sample ( Figure 4 a and 4b). It can be modelled with the stretched -\nexponential (SE) time dependence \n \nSE etPt , (1) \nwell suited for dynamical local fields in diluted spin systems39. Here denotes the muon \nrelaxation rate and the stretch exponent. For each sample the relaxation rate is found to \nincrease notably with increasing temperature (P(t) curves in Figure 4 a and 4b decrease faster \nat higher temperature ). This increase of the relaxation rate is very similar in the annealed and \nnon-annealed sample s for a given concentration of dopants . Therefore, the muon relaxation in \nthe Fe -doped BaTiO 3- system is attributed to the fluctuating paramagnetic moments . \n \nDirect confirmation of this statement is provided by comparing the ZF data and the data \nobtained in the weak longitudinal field (LF) of 1 mT. Such field comparison shows a very \nsimilar relaxation response (Figure 4c). We note that in addition to the dynamical model of \nequati on (1), the Lorentzian Kubo -Toyabe (KT) model38 \n11 \n \n\n KT121e33tP T t , (2) \nalso provides a satisfactory fit to the ZF data. The KT model is applicable to disordered static \nlocal fields with the distribution width = 7.0∙10-2 mT = 2∙135.5 MHz/T is the muon \ngyromagnetic ratio However, in this static case the 1 mT longitudinal field would almost \ncompletely remove the relaxation of P(t), which clearly contradicts the experiment (Figure \n4c). The muon relaxation due to local fields fluctuating with the exchange frequency \n\neB ~1 THz kJ\n on the other hand, is not affected by such a small LF field38, because\ne 1 mT=0.85 MHz\n. \n \n \nFigure 4: SR characterization of 6 H-BaTiO 3-. The temperature dependence of muon \npolarization in (a) 10% and (b) 20% Fe -doped 6 H-BaTiO 3- samples in zero applied field \nshowing increased relaxation rates with increasing temperature . The upper panels correspond \nto non -annealed samples and the lower panels to annealed samples. The solid lines are fits to \nthe stretched expon ential (SE) model of e quation (1 ). (c) The room -temperature muon \npolarization of the F20BTOa sample in ZF and in a weak longitudinal applied field. The dashed \nlines correspond to the static Kubo -Toyabe (KT) model of e quation (2 ), while the solid lines \nare fits to the model encompassing both static KT and dynamical SE relaxation (see text for \ndetails ). The error bars of muon polarization data are defined as a square root of the total \nnumber of detected positrons. \n \n0.10.40.71.0\n0 2 4 6 810 120.20.40.60.81.0\n0 2 4 6 810 120.10.40.71.00.40.71.0\n0 2 4 6 810 120.40.71.0323 K\n423 K\n523 K623 K\n723 K\n823 Ka Muon polarizationF10BTOc\nF20BTOa\nt (s)0 mT\n data\n KT\n KT * SE\n1 mT\ndata\n KT\n KT * SE \n293 K\nF10BTOa 323 K\n423 K\n523 K623 K\n723 K\n823 K \nt (s)323 K\n423 K\n523 K623 K\n723 K\n823 Kb \nF20BTO\nF20BTOa 323 K\n423 K\n523 K623 K\n723 K\n823 K \nt (s) \n12 \n A close r look at the ZF and LF data of the F20BTOa sample at room temperature, however, \nreveals a somewhat decreased muon relaxation in the LF experiment (Figure 4 c). This can be \nattributed to the presence of some weaker static relaxation in addition to the dominant \ndynamical one. A simultaneous fit of both datasets to the corresponding model \n SE KT P t P t P t\n yields the static -field distribution width of = 6.1 (3)∙10-3 mT. This value \nrepresents the upper bound of the frozen -FM-moments ’ contribution in the annealed samples , \nsince the static -field distribution may also partially arise from nuclear magnetic moments . \nHowever, for homogeneously distributed FM moments the calculated field -distribution widths \nat both oxygen sites , in the vicinity of which the muons are highly likely to reside in oxides38, \nare almost an order of magnitude larger , = 3.4 (2)∙10-2 mT (see Methods) , even for the \nvery small average value of the FM moments in F20BTOa (1.5∙10-3B at 300 K). As the \ninteraction strength between the muon and the FM moment decreases with the cube of their \ndistance, t he much smaller experimental value of can be reconciled by considering \nclustering of the FM moments . This thus provides microscopic evidence that , in contrast to the \nhomogeneous distribution of PM moments evidenced by ESR , the FM moments are not \nuniformly distributed throughout the sample . \nDiscussion \nThe small size of the average FM moments and their non -uniform distribution in the annealed \nFe-doped 6H-BaTiO 3- samples cast doubt on the intrinsic origin of these moments . It is \ntherefore crucial to critically assess th e origin of ferromagnetism in this system . It could well \nbe that the FM behaviour either correspond s to some kind of “impurity” phase or that the \nspinodal decomposition occurs and very small dispersed lattice -matched FM clusters of Fe3+ \nions form during the annealing process. The latter scenario has been recently demonstrated17 \nin Mn -doped SrTiO 3. In ord er to check for the presence of such small clusters that would \nminimally deform the local crystal structure and would be practically undetectable even by high -\nresolution imaging techniques, we performed a magnetization relaxation experiment on the FM \nF20BT Oa sample. After being exposed to the magnetic field of 1 T for 20 minutes, the \nsubsequent remnant relaxation in zero field was measured. No magnetization relaxation \n13 \n towards zero has been detected on the time scale of several minutes (Figure 2c). This implies \nan unexpectedly slow relaxation process within the Stoner -Wohlfarth model of single -domain \nFM (nano)particles40 and thus excludes the presence of small FM clusters of Fe3+ ions. \nTherefore, the alternative scenario of “impurity” ferromagnetism is much more plausible and \nshould be carefully considered. The Curie temperature s of the most common iron oxides41 that \nmay develop as a by-product during the synthesis are much higher than Tc1 ~ 700 K; ~950 K \n(Fe2O3) and ~850 K (Fe 3O4). On the other hand, the Curie temperature of ~680 K has been \nrecently reported for the 3C bulk BaTi0.95Fe0.05O3- sample, where a much larger magnetic \nmoment of 0 .75B per Fe has been found42. The high -temperature ferromagnetism has regularly \nbeen encountered in this pseudocubic polymorph, even at much higher Fe -doping \nconcentrations (up to 75%), where the 3 C phase is stabilized only by the confined geometry of \nthin films and nanoparticles43–46. We note though that the Curie temperature of an impurity \nphase is usually size dependent for nano -particles, as well as it may depend on the level of \ndoping. Therefore, the fair agreement of Tc1 with the Curie temperature of the 3C bulk \nBaTi 0.95Fe0.05O3- sample cannot be taken as a solid proof that the FM cha racter of the Fe -doped \n6H-BaTiO 3- is due to the impurity 3 C phase . However, an unambiguous evidence that this is \nindeed the case comes from the slightly doping dependent Tc3 transition temperature (\n10%\nc3 430 K T\n,\n20%\nc3 450 K T ; see Figure 3d), which exactly coincides with the doping dependent \ncubic-to-tetragonal structural transition of the 3 C phase43. Moreover, t his structural transition \nis clearly reflect ed in the FM ESR line-shape change from the high-temperature single line to a \nmulti-feature spectrum below Tc3, which corroborates the change of the local sym metry from \nbeing highly symmetric above Tc3 to being less symmetric below Tc3. It is also worth noting that \nin epit axially grown thin films, the 3C phase grows at the initial stages, while later on a \ndisordered phase composed of 3C and 6 H intergrowths forms43. Since increasing thickness \ndecreases the FM response of thin -film samples47, their ferromagnetism is most naturally \nattributed to the 3C phase, for which it is also theoretically pre dicted as the ground state42. \nAll the above -presented arguments give clear evidence that the FM behaviour of the annealed \nFe-doped BaTiO 3- samples is related to internal strains that locally destabilize the 6 H phase in \n14 \n favour of the 3 C phase , much as it happens globally in the confined geometries of thin films \nand nanoparticles . The SR results yielding only minor differences between the FM annealed \nand PM non-annealed samples , as well as the non-uniform distrib ution of FM moments , \ncorroborate this scenario. Furtherm ore, ferromagnetism arising from the sparse strain -\nstabilized defect regions with the 3C crystal structure explain s both the extremely small \nmagnitude of the average ordered Fe moment and its unusual decreas e with increasing doping \nconcentration; the increasing doping level being known to destabilize the 3C phase35,48. Finally, \nin the non -annealed samples that were hea t-treated at lower temperature (see Methods for \ndetails ) the crystallites are under significantly less strain than in the densely sintered annealed \nsamples after additional annealing at higher temperature . Therefore, i n the latter samples \nstrong strain fields along crystallographically mismatch ed grain boundaries , or around plane \ndefect s, are likely regions for nucleation of stable pseudocubic domains with their own intrinsic \nferromagnetism. These are not single -domain nano -sized FM defects, as m agnetization \nrelaxation is absent. \nOur comprehensive magnetic investigation of various Fe-doped 6 H-BaTiO 3- samples has thus \nrevealed that the FM response observed in the annealed samples is not intrinsic to the \nhexagonal crystallographic phase, as broadly speculated before. This conclusion is based on our \ncomplementary bulk and local -probe magnetic investigation s, providing new microscopic insight \nand extending the experimental temperature range far beyond all previous reports. Bulk \nmagnetization measureme nts have demonstrated that extremely small average static magnetic \nmoments develop in the annealed samples below Tc1 ~ 700 K and several additional magnetic \ninstabilities occur a s the temperature is lowered further . Below the Tc3 transition (at aroun d 430 \nK and 450 K in 10% - and 20% -doped samples, respectively ) a signi ficant symmetry reduction \nof the local structure is detected by ESR. The Tc1 and Tc3 transitions coincide with the \nferromagnetic and ferroelectric transitions of the Fe -doped 3 C-BaTiO 3 polymorph, respectively, \nwhile the microscopic characteristics determined by SR imply non-uniformly distributed FM \nregions in the samples . This demonstrates that the FM response of the Fe-doped 6 H-BaTiO 3- \nsystem originates from sparse regions where the pseudocubic structural poly morph is stabilized \nby strain fields. The strain -induced local competition between the two structural (and magnetic) \n15 \n phases is inherent only to den sely sintered annealed samples. Such scenario of ferromagnetism \nmay turn to be important for other DMO materials, where the competi tion between different \nstructural phases is intrinsically present , in particular in confined geometries where strain \neffects are enhanced . \nMethods \nSamples \nHigh quality 6 H-BaTiO 3- polycrystalline samples with c = 10, 20% Fe3+ ions substituted for Ti4+ \nwere synthesized according to the procedure thoroughly explained in Ref. 28. For each \ncomposition, after heat treatment at 1250 °C, a part of the sample was additionally annealed \nin oxygen atmosphere at 1500 °C, typically for 5 – 10 hours. We label the non -annealed and \nannealed samples with F cBTO and F cBTOa, respectively. The variations of the annealing time \nshowed no significant influence on the magnetic properties28. X-ray powder diffraction (XRD) \nwas used to verify the single -phase hexagonal struct ure of all our samples. The elemental \nanalysis was performed by Energy Dispersion X -Ray Spectroscopy (EDX) on polished ceramic \nsurfaces with JSM -7100 F (Jeol) field -emission scanning ele ctron microscope equipped with an \nx-ray detector (X -Max 80, Oxford Instrument). Ten EDX characterizations were performed on \neach sample and were statistically treated to obtain average values and standard deviations. \nThe iron concentration in the nominally 10%-doped samples was found at 10.3±0.5%, while in \n20%-doped samples it amounted to 19.9±0.6%. For all our samples t he analytically determined \ncompositions thus correspond to the nominal compositions within the small error bars. \nBulk magnetization \nBulk ma gnetization measurements were performed on a couple of Quantum Design SQUID \nmagnetometers and a vibrating sample magnetometer (VSM) as a function of temperature \n(between 2 and 850 K) in various magnetic fields and at various fixed temperatures as a \nfunctio n of varying magnetic field. Zero -field-cooled (ZFC) and field -cooled (FC) temperature -\ndependent data were collected. A high -temperature insert was used for measurements above \n400 K. The SQUID measurements were performed on samples sealed in quartz capilla ries, while \n16 \n boronitride sample holders were used with VSM. The signal from the sample holders was \ncarefully evaluated. The size of the ordered magnetic moment was estimated from the \nsaturation value of the FM component in the hysteresis, by sub tracting the linear ly increasing \nPM part. \nESR \nElectron spin resonance was measured with home -built resonator -cavity based X -band (9.4 \nGHz) and K -band (24.3 GHz) spectrometer s. In the X band, a continuous -flow cryostat was \nused in the temperature range 5 – 300 K , while a preheated -nitrogen -flow heating system was \nused above room temperature up to 620 K. The samples were sealed in ESR silent quartz tubes. \nThe ESR intensity was calibrated at room temperature with a reference sample (CuSO 4∙5H2O). \nThe ESR susceptibil ity of the non -annealed PM samples ESR = 0.06(3) Am2/ T per mole Fe was \nfound compatible with the bulk molar susceptibility b = 0.086 Am2/T, proving that ESR was \ndetect ing the intrinsic signal of Fe dopants in BaTiO 3. \nThe second moment36 of the ESR line M2 arising from dipolar interactions is given by \n \n22 2\n40\n2B 63cos 1 3144jk\nk jkM S S gr (3) \nwhere 0 is the vacuum permeability. The sum runs over all neighbours k of a given site j, \nconnected by the vector rjk. jk is the angle between rjk and the applied magnetic field . M2 was \ncalculated by powder -averaging e quation (3) for a lattice fully occupied with Fe3+ ions on either \nthe Ti(1) or Ti(2) site. In the limit of negligible exchange interactions M2 yields the ESR peak-\nto-peak line width \n 2B3 B M g \n for the fully occupied lattice and the line width \n B c c B\n in the dilute limit17,36. We find \n20%\nTi(1)23 mT B and \n20%\nTi(2)31 mT B , which is far \nbelow the experimental values . In the exchange narrowing limit (\n\nBBk J g B ), on the other \nhand, the ESR line width is given by \n \n2 B B B M g k J . The second moment M2 is quadratic in the \nanisotropic interaction (i.e., in the doping level c), therefore B also scales linearly wi th c for a \nlinearly dependent J. \n17 \n \nSR \nIn a SR experiment almost 100% spin -polarized muons are implanted into a sample. As \nparticles possessing a magnetic moment, they interact with local magnetic fields, which drive \nthe time dependence of muon polarization P(t). At the time of the muon decay a positron is \nemitted preferentially in the direction of the muon ’s spin direction . The detection of positrons \ntherefore allows for the reconstruction of P(t) and thus for the determination of a magnitude, \ndistribution, and fluctuation rate of local internal magnetic fields . \nMuon spin relaxation measurement were performed on the EMU instrument at the ISIS facility, \nRutherford Appleton Laboratory, UK. A flat -plate furnace was used to cover the temperature \nrange between 290 and 820 K. The background signal from a titanium sample holder was \nestimated with a hematite reference sample. It was found to typically amount to ~30% of the \ntotal signal and was subtracted from the SR data shown above. The measurements were \ncarried out with detectors grouped in the forward -backward direction with respect to the muon \nbeam in zero field (ZF) and in longitudinal applied magnetic field (LF), while calibration \nmeasurements were performed in a weak transverse field of 2 mT. \nThe dipolar -magnet ic-field distribution at a given muon stopping site , \n\n01d(B)dNDNB , can be \naccurately modelled , as demonstrated in Ref. 49. In 6 H-BaTiO 3- it was calculated by randomly \npopulating fraction c of all Ti sites with frozen Fe3+ moments . The magnetic field at a chosen \ncrystallographic site was the n calculated in N0 = 2,048 different crystallographic cells. For each \npoint al l magnetic moments within a spherical region around it large enough to ensure \nconvergence were taken into account . \n \n1. Dietl, T. Dilute magnetic semiconductors: Functional ferromagnets. Nat. Mater. 2, 646–\n648 (2003). \n18 \n 2. Kuroda, S. et al. Origin and control of high -temperature ferromagnetism in \nsemiconductors. Nat. Mater. 6, 440–446 (2007). \n3. Dietl, T. A ten -year perspective on dilute magnetic semiconductors and oxides. Nat. \nMater. 9, 965–974 (2010). \n4. Yamada, Y. et al. Electrically induced ferromagnetism at room temperature in cobalt-\ndoped titanium dioxide. Science 332, 1065–1067 (2011). \n5. Chen, G. et al. Resistive switching and magnetic modulation in cobalt-doped ZnO. Adv. \nMater. 24, 3515–3520 (2012). \n6. Pellicer, E. et al. Nanocasting of mesoporous In -TM (TM = Co, Fe, Mn) oxides: Towards \n3D diluted -oxide magnetic semiconductor architectures. Adv. Funct. Mater. 23, 900–911 (2013). \n7. Coey, J. M. D. & Chambers, S. A. Oxide dilute magnetic semiconductors —fact or fiction? \nMRS Bull. 33, 1053–1058 (2008). \n8. Izyumskaya, N., Alivov, Y. & Morkoç, H. Oxides, oxides, and more oxides: High -κ oxides, \nferroelectrics, ferromagnetics, and multiferroics. Crit. Rev. Solid State Mater. Sci. 34, 89–179 (2009). \n9. Pulizzi, F. & Chambers, S. Is it really intrinsic ferromagnetism? Nat. Mater. 9, 956–957 \n(2010). \n10. Garcia, M. A. et al. Sources of experimental errors in the observation of nanoscale \nmagnetism. J. Appl. Phys. 105, 013925 (2009). \n11. Valant, M., Kolodiazhnyi, T., Axelsson, A. -K., Babu, G. S. & Alford, N. M. Spin ordering \nin Mn -doped KTaO 3? Chem. Mater. 22, 1952–1954 (2010) . \n12. Ogale, S. B. Dilute doping, defects, and ferromagnetism in metal oxide systems. Adv. \nMater. 22, 3125–3155 (2010). \n19 \n 13. Sawicki, M., Stefanowicz, M., Ney, A., Sensitive SQUID magnetometry for studying \nnanomagnetism. Semicond. Sci. Technol. 26, 064006 (2011). \n14. Valant, M. et al. The origin of magnetism in Mn -doped SrTiO 3. Adv. Funct. Mater. 22, \n2114–2122 (2012). \n15. Bonanni, A., Dietl, T. A story of high -temperature ferromagnetism in semiconductors. \nChem. Soc. Rev . 39, 528 -539 (2010). \n16. Bonanni, A. et al. Controlled aggregation of magnetic ions in a semiconductor: an \nexperimental demonstration. Phys. Rev. Lett. 101, 135502 (2008). \n17. Zorko, A., Pregelj, M., Luetkens, H., Axelsson, A. -K. & Valant, M. Intrinsic \nparamagnetism and aggregation of manga nese dopants in SrTiO 3; Phys. Rev. B 89, 094418 (2014). \n18. Ray, S. et al. High temperature ferromagnetism in single crystalline dilute Fe -doped \nBaTiO 3. Phys. Rev. B 77, 104416 (2008). \n19. Burbank, R. D. & Evans, H. T. The crystal structure of hexagonal barium titanate. Acta \nCrystallogr. 1, 330–336 (1948). \n20. Keith, G. M., Sarma, K., Alford, N. M. & Sinclair, D. C. Electrical properties of 6H -\nBaTi 0.95M0.05O3−δ ceramics where M = Mn, Fe, Co and Ni . J. Elect roceramics 13, 305–309 (2004). \n21. Wei, X. K. et al. Origin of ferromagnetism and oxygen -vacancy ordering induced cross -\ncontrolled magnetoelectric effects at room temperature. J. Appl. Phys. 111, 073904 (2012). \n22. Chakraborty, T., Ray, S. & Itoh, M. Defect -induced magnetism: Test of dilute magnetism \nin Fe-doped hexagonal BaTiO 3 single crystals. Phys. Rev. B 83, 144407 (2011). \n23. Wei, X. K. et al. Structure, electrical and magnetic property investigations on dense Fe -\ndoped hexagonal BaTiO 3. J. Appl. P hys. 110, 114112 (2011). \n20 \n 24. Lin, F., Jiang, D., Ma, X. & Shi, W. Influence of doping concentration on room -\ntemperature ferromagnetism for Fe -doped BaTiO 3 ceramics. J. Magn. Magn. Mater. 320, 691–694 \n(2008). \n25. Lin, F., Jiang, D., Ma, X. & Shi, W. Effect of annealing atmosphere on magnetism for Fe -\ndoped BaTiO 3 ceramic. Phys. B 403, 2525–2529 (2008). \n26. Lin, F. & Shi, W. Effects of doping site and pre -sintering time on microstructure and \nmagnetic properties of Fe -doped BaTiO 3 ceramics. Phys. B 407, 451–456 (2012). \n27. Coey, J. M. D., Venkatesan, M. & Fitzgerald, C. B. Donor impurity band exchange in \ndilute ferromagnetic oxides. Nat. Mater. 4, 173–179 (2005). \n28. Valant, M., Arčon, I., Mikulska, I. & Lisjak, D. Cation order–disorder transition in Fe -\ndoped 6H -BaTiO 3 for dilute room-temperature ferromagnetism. Chem. Mater. 25, 3544–3550 (2013). \n29. Chikada, S., Hirose, K. & Yamamoto, T. Analysis of local environment of Fe ions in \nhexagonal BaTiO 3. Jpn. J. Appl. Phys. 49, 091502 (2010). \n30. Grey, I. E., Li, C., Cranswick, L. M. D., Roth, R. S. & Vanderah, T. A. Structure Analysis \nof the 6H –Ba(Ti, Fe3+, Fe4+)O3- solid solution . J. Solid State Chem. 135, 312–321 (1998). \n31. Chakraborty, T., Meneghini, C., Aquilanti, G. & Ray, S. Microscopic di stribution of metal \ndopants and an ion vacancies in Fe -doped BaTiO 3−δ single crystals. J. Phys. Condens. Matter 25, 236002 \n(2013). \n32. Gaj, J. A., Planel, R. & Fishman, G. Relation of magneto -optical properties of free \nexcitons to spin alignment of Mn2+ ions in Cd 1-xMnxTe. Solid State Commun . 29, 435-438 (1979). \n33. Ohi, K., Arai, H., Ishige, T. & Shimokoshi, M. ESR study of phas e transition in hexagonal \nBaTiO 3. J. Phys. Soc. Jpn. 58, 3781–3787 (1989). \n34. Shimokoshi, M. & Ohi, K. ESR study of ferroelect ric phase transition in h -BaTiO 3. J. Phys. \nSoc. Jpn. 59, 3629–3634 (1990). \n21 \n 35. Böttcher, R., Langhammer, H. T., Müller, T. & Abicht, H. -P. 3C–6H phase transition in \nBaTiO 3 induced by Fe ions: an electron paramagnetic resonance study. J. Phys. Condens. Matter 20, \n505209 (2008). \n36. Bencini, A. & Gatteschi, D. EPR of Exchange Coupled Systems . (Springer -Verlag, 1990). \n37. Kittel, C. On the theory of ferromagnetic resonance absorption. Phys. Rev. 73, 155–161 \n(1948). \n38. Yaouanc, A. & Dalmas de Réotier, P. Muon Spin Rotation, Relaxation and Resonance . \n(Oxford University Press, 2011). \n39. Uemura, Y. J., Yamazaki, T., Harshman, D. R., Senba, M. & Ansaldo, E. J. Muon -spin \nrelaxation in AuFe and CuMn spin glasses. Phys. Rev. B 31, 546–563 (1985). \n40. Chuev, M. A . & Hesse, J. Nanomagnetism: extension of the Stoner –Wohlfarth model \nwithin Néel’s ideas and useful plots. J. Phys. Condens. Matter 19, 506201 (2007). \n41. Cornell, R. M. & Schwertmann, U. The Iron Oxides: Structure, Properties, Reactions, \nOccurrences and U ses. (Wiley -VCH, 2003). \n42. Xu, B. et al. Room -temperature ferromagnetism and ferroelectricity in Fe -doped BaTiO 3. \nPhys. Rev. B 79, 134109 (2009). \n43. Maier, R., Cohn, J. L., Neumeier, J. J. & Bendersky, L. A. Ferroelectricity and \nferrimagn etism in iron -doped BaTiO 3. Appl. Phys. Lett. 78, 2536 (2001). \n44. Matsui, T., Taketani, E., Fujimura, N., Ito, T. & Morii, K. Magnetic properties of highly \nresistive BaFeO 3 thin films e pitaxially grown on SrTiO 3 single -crystal substrates. J. Appl. Phys. 93, 6993 \n(2003). \n45. Rajamani, A., Dionne, G. F., Bono, D. & Ross, C. A. Faraday rotation, ferromagnetism, \nand optical pro perties in Fe -doped BaTiO 3. J. Appl. Phys. 98, 063907 (2005). \n22 \n 46. Yang, L. et al. Magnetic properties of BaTiO 3 and BaTi 1−xMxO3 (M=Co, Fe) nanocrystals \nby hydrothermal method. J. Magn. Magn. Mater. 350, 1–5 (2014). \n47. Maier, R. & Cohn, J. L. Ferroelectric and ferrimagnetic iron -doped thin -film BaTiO 3: \nInfluence of iron on physical properties. J. Appl. Phys. 92, 5429 (2002). \n48. Nguyen, H. M. et al. Tetragonal and hexagonal polymorphs of BaTi 1−xFexO3−δ \nmultiferroics using x -ray and Raman analyses. Appl. Phys. Lett. 99, 202501 (2011). \n49. Zorko, A., Adamopoulos, O., Komelj, M., Arčon, D. & Lappas, A. Frustration -induced \nnanometre -scale inhomogeneity in a t riangular antiferromagnet. Nat. Commun. 5, 3222 (2014). \nAcknowledgements \nWe acknowledge the financial support of the Slovenian Research Agency Programs P1 -0125, \nP2-0377, P1 -0112 and Projects N2 -0005, Z1 -5443, as well as the Research Support \nProgramme of University of Zagreb . The SR study has been supported by the European \nCommission under the 7th Framework Programme through the 'Research Infrastructures' \naction of the 'Capacities' Programme, Contract No. 283883 -NMI3 -II. \nAuthor contributions \nA.Z. and M.V. designed and supervised the project. I.A., I.M. and M.V. synthesized and \nstructurally characterized the samples. Z.J. and D.P. performed bulk magnetization \nmeasurements. M.P. and M.G. conducted ESR measurements. A.Z., M.G. and M.T. carried out \nthe SR study. All the experimental data were analysed by A.Z., who also wrote the paper. All \nauthors discussed the results and reviewed the manuscript. \nAdditional information \nCompeting financial interests \nThe authors declare no competing financial interests. " }, { "title": "1809.08644v1.Ferromagnetic_resonance_in_thin_ferromagnetic_film_with_surface_anisotropy.pdf", "content": "Ferromagnetic resonance in thin ferromagn etic film with surface anisotropy \n \nN. A. Usov1,2\n \n1National University of Science and Tec hnology «MISiS», 119049, Moscow, Russia \n2Pushkov Institute of Terrestrial Magnetism, Ionosphere and Radio Wave Propagation, \nRussian Academy of Sciences, IZMIRAN, 108480, Troitsk, Moscow, Russia \n \nAbstract The ferromagnetic resonance frequencies are obtaine d for a thin ferromagnetic film with surface \nanisotropy for the cases when th e external magnetic field is applied perpe ndicularly or parallel to the film surface, \nand for various combinations of boundary conditions on the film surface. It is shown that in the presence of surface \nanisotropy the ferromagnetic resonance frequency essentially depends both on th e film thickness and on the value of \nthe surface anisotropy constant. The results obtained provide a basis for the correct interpretation of experimental \ndata obtained by means of broadband ferroma gnetic resonance in thin film structures. \n \nPACS: 75.50.Kj, 75.80.+q, 75.60.Ej, 75.60.Jk \nKeywords: Thin ferromagnetic film, surface magnetic anisotropy, ferromagnetic resonance frequency \n \nI. Introduction \nThin ferromagnetic films with su rface magnetic anisotropy are cu rrently of great technological \ninterest [1-4]. In particular, the ultra-thin CoFe B/MgO bilayers [2-4] exhibit a perpendicular magnetic \nanisotropy. The latter provides a high thermal stability and is highly de sirable for reducing the critical \ncurrent for spin-torque switching [2]. Therefore, thin CoFeB ferromagnetic films are considered currently \nas promising electrode material in magnetic rando m access memory and spin-transfer torque memory \ndevices. \n From physical considerations, it is obvious that only the surface magnetic anisotropy [5-8] is \ncapable of providing an out-of-plane magnetization in a thin ferromagnetic film made of a soft magnetic \nmaterial. Actually, if a surface anisotropy constant is negative and sufficiently large in absolute value, the \norientation of the unit magnetization v ector perpendicular to the film pl ane is energetically favorable, in \nspite of a significant increase in the magnetostatic en ergy of the film. Experimentally, a rotation of the \nunit magnetization vector as a function of the film thickness was observed in a number of experiments \nwith thin ferromagnetic films of iron, cobalt, and other ferromagnets [9-12]. \nThe existence of surface magnetic anisotropy may be related to various physical origins, such as \nthe specifics of the spin-orbit interaction on the ferromagnetic surface [13], the difference in atomic periods of various layers of a multilayer sample, or a film substrate [14], th e presence of non-uniform \nmechanical stresses near the inte rface [15], etc. However, in terms of phenomenological description of \nmagnetic phenomena [5], it is important that the energy density of surface anisotropy is concentrated in a \nvery narrow region near the sample surface. Therefor e, it is a surface contri bution to the total energy \nwhich is proportional to the surface area of the film , and not to its volume. The general variational \napproach [5] shows that the influence of surface ma gnetic anisotropy on the behavior of a ferromagnet is \nmanifested only under special boundary condition. The latt er acts on the sample surface or the interface \nbetween different materials [5,16]. \nSince the theoretical calculation of the surface anisotropy constant Ks is rather complicated [13-\n15], it is desirable to have reliable methods to dete rmine this phenomenological constant experimentally. \nOne of such methods can be the ferromagnetic res onance (FMR) [17,18], in whic h the film magnetization \nundergoes a rapid precession near a certain equilib rium position of the unit magnetization vector. The \nbroadband FMR [19-23] can provide valuable information on the presence or absence of the energy \ninteractions on the surface of a ferromagnetic film, or on the interfaces of ferromagnetic multilayers. The \ninfluence of surface anisotropy on the ferromagnetic resonance in a thin ferromagnetic film was previously considered in seminal papers by Rado and Weertman [24,25]. Un fortunately, no explicit \n 1formulas for the FMR frequency were actually obt ained. Whereas in the present paper, the FMR \nfrequencies have been derived for ferromagnetic f ilm with surface anisotropy for the cases when the \nexternal magnetic field is applied perpendicularly or parallel to the film surface. Besides, various \ncombinations of boundary conditions on the film surface s have been studied. Thes e results seem helpful \nfor correct interpretation of the experimental data obt ained by means of the broadband FMR in thin film \nstructures. \n \nII. Out-of-plane external magnetic field \n Consider a thin ferromagnetic film of thickness L parallel to the XY plane and located in the \ndomain 0 < z < L. The unit magnetization vector of the film ()t r,rrα satisfies the Landau-Lifshitz-Gilbert \n(LLG) equation \n[]⎥⎦⎤\n⎢⎣⎡\n∂∂× + × − =∂∂\ntHtefαα κ α γαrrrrr\n, ( 1 ) \nwhere γ is the gyromagnetic ratio, κ being the phenomenological dampi ng constant. The total effective \nmagnetic field of the film efHr takes into account the contributions due to the exchange and dipolar \ninteractions as well as from th e magnetic anisotropy energy. \n \n0H HMw\nMCH\nsa\nsefr r\nrr r+′+∂∂− Δ =αα . ( 2 ) \nHere C = 2A is the exchange constant, Ms is the saturation magnetization, is the vector of the \nhomogeneous applied magnetic field, 0Hr\nH′r is the vector of the demagnetiz ing field which is created by the \nvolume and surface magnetic charges distributed in the volume and on the surface of the film, \nrespectively. Finally, ()αrw is the energy density of the magne tic anisotropy in the film volume. \n The presence of surface magnetic anisotropy can be described [5-8] by introducing a surface \ninteraction with the ener gy density per unit area \n \n ()2n K ws sarrα= . ( 3 ) \nHere Ks is the phenomenological surface anisotropy constant, and nr is the unit vector of the external \nnormal to the film surface. In th e presence of the surface magnetic anisotropy, Eq. (3), the boundary \ncondition for the unit magnetiza tion vector is given by [5] \n \n () ()(α α α)α rrrr rrr\nn n n KnCs− =∂∂2 . ( 4 ) \nIt works on the upper and lower surfaces of the ferro magnetic film. In the absence of surface anisotropy \non one or both surfaces of the film, Ks = 0, the usual boundary condition, 0= ∂ ∂nαr, acts on the \ncorresponding surface. \n Suppose that the ferromagnetic film is placed in a sufficiently strong external magnetic field \nperpendicular to the film surface, ()0 0 , 0 , 0H H=r. In what follows we neglect for simplicity the effect of a \nsmall volume magnetic anisotropy of the film, setting ()0=αrw . In equilibrium, in the absence of \nmagnetization perturbations, the unit magnetization vector is perpe ndicular to the film surface, \n. Indeed, in the case considered the v ector of effective ma gnetic field has only z component, ()(1 , 0 , 00=αr)\n) (s ef M H H H H π4 , 0 , 00) 0 (\n0) 0 (− = ′+ =rrr. Therefore, the equilibrium condition for the unperturbed unit \nmagnetization vector \n . () ()[]00 0= ×efHrrα . ( 5 ) \n 2as well as the boundary condi tion (4) are satisfied. \n Let us now consider small deviations of th e unit magnetization vector from the equilibrium \nposition. The first-order correction of the perturbation theory [5] to the unit magnetization vector is given \nby () () ()()0 , ,1 1 1\ny xα α α=r. The components of the first-order co rrection satisfy the equation of motion \n \n ()\n() ()[]() ()[]()()\n⎥⎦⎤\n⎢⎣⎡\n∂∂× + × − × − =∂∂\ntH Htef ef1\n0 1 0 0 11αα κ α γ α γαrrrrrrr\n (6) \nSince the magnetization in the film pl ane is homogeneous, it is reasonable to assume that the perturbation \nof the magnetization depends only on the z coordinate. Then the first order correction to the effective \nmagnetic field vector is given only by the exchange interaction, so that \n \n () ()\n⎟⎟\n⎠⎞\n⎜⎜\n⎝⎛\n∂∂\n∂∂= 0 , ,21 2\n21 2\n) 1 (\nz MC\nz MCHy\nsx\nsefα α r. ( 7 ) \nTaking into account Eqs. (6), (7) and neglec ting the attenuation, κ → 0, one obtains the following \nequations of motion for the component s of the unit magnetization vector \n \n () ()\n()()1\n21 2 1\ny H my\nsx\nz MC\ntα ω ωαγα− +∂∂=∂∂; ( 8 a ) \n () ()\n()()1\n21 2 1\nx H mx\nsy\nz MC\ntα ω ωαγα− −∂∂− =∂∂, ( 8 b ) \nwhere we denote ωH = γH0 and ωm = 4πγMs. The boundary conditions to Eq. (8) follow from Eq. (4) \n \n ()()()()t L Kzt LCx sx, 2,1\n11\nαα− =∂∂; ()()()(t L Kzt LCy sy, 2,1\n11\nαα− =∂∂); (9a) \n ()()()()t KztCx sx, 0 2, 01\n21\nαα=∂∂; ()()()(t KztCy sy, 0 2, 01\n21\nαα=∂∂)\n). (9b) \nHere it is assumed that the surface anisotropy constants at the upper, K1s, and lower, K2s, film surfaces can \nbe different. \n If one of the surface anisotropy constants is zero, for example, K2s = 0, the solution of Eq. (8), (9) \nhas the form of the unit ma gnetization vector precession \n \n ; ()() ( )( t kz A t zx ω α cos cos ,1=()()()()t kz B t zy ω α sin cos ,1= . (10) \nFrom the equations of motion (8) one obtains for the precession frequency \n \n 2kMC\nsm Hγ ω ω ω+ − = , ( 1 1 ) \nwhereas the boundary conditions (9) lead to the dispersion equation \n \n ()CKkL ks12tan= . ( 1 2 ) \nIn the limit kL << 1 for the lowest quasi-homogeneous pre cession mode one obtains from Eq. (12) \n 3 CLKks12= ; L MK\nss\nm H12γ ω ω ω+ − = . (13) \nFor inhomogeneous FMR modes in the limit 11<1 kOe) are much\nhigher than the SAF saturation field [ Hs\u0019300 Oe; Fig. 1(c)].\nThe resonance field, Hr, of each resonance line show a pro-\nnounced angular dependence, as seen in Fig. 2(b). All lines\nshow bi-axial anisotropy with, however, opposite hard and\neasy axes for L mand L e1;e2. While L mexhibits easy axes\n(minimum Hr) at 45\u000ewith respect to the main xandyaxes\nof the array [Fig. 1(a)], the edge-modes, L e1and L e2, have\neasy axes along xandy. Such anisotropic behavior was ex-\nplained in Refs. 9 and 34 by the inter-element dipole coupling\nleading to a non-uniform effective field distribution within the\nindividual discs, even at rather high external fields.\nIn order to explain the origin and behavior of the three FMR\nmodes observed, we performed detailed micromagnetic sim-\nulations [35], using MuMax3 [36, 37]. The layout of each\nelement as well as the base diameter ( d0) used were those\nobserved experimentally [Fig. 1(c)]. Figure 2(c) displays the\ncalculated spectra for two isolated single-layer Py discs (7 nm\nthick) with diameters d0=150 nm (blue line) and d=0:85d0\n(red dashed line). This difference in size corresponds to the\nSEM-measured difference in the diameters of the bottom and\ntop layers of our SAF nanodiscs; cf. Fig. 1(b). Both simu-\nlated spectra show two resonance lines corresponding to the\nmain and edge-resonance modes, illustrated by the insets to\nFig. 2(c). Of note is that despite line L mis associated with the\nuniform mode of a continuous Py film, the related excitation\nwithin a nanodisc is not fully spatially uniform, which is why\nwe call it main rather than uniform FMR mode. Importantly,\nneither spectrum shows any trace of a third line observed on\nthe experiment, nor a related additional line can be obtained\nwhen the two spectra are superposed.\nThe simulated spectra are substantially different when the\ntwo discs of d0=150 nm and d=0:85d0form a SAF ele-\nment, as shown in Fig. 2(d). Three resonance lines are visible,\nwith the positions and intensities resembling the experimen-\ntal data. The insets to Fig. 2(d) clarify why we associate the\nhigh-field lines with edge-mode resonances. Whereas the L e1\nmode is excited mostly in the bottom layer, the L e2mode is\nmore intensive in the top layer. Importantly, when the top and\nbottom layers are of equal diameter, only two modes, L mand\nLe1, are excited; the third mode appears only when the top and\nbottom diameters become noticeably different (by 5–10 %).\nFigure 2(e) shows the simulated angle-dependence of the\nresonance fields of the L m, Le1, and L e2modes obtained for\nan array of SAF nanodiscs. The simulations are in excellent\nagreement with the experimental resonance-field versus angle\ndata of Fig. 2(b), which further validates our interpretation of\nthe observed modes and their properties.\nFigure 3(a) compares the SAF FMR spectra measured as3\nFIG. 2. (a) FMR spectra measured at RT for 150-nm SAF array at 0, 45, and 90\u000e-oriented external field. L m, Le1and L e2mark resonance\nlines attributed to main and two edge-modes, respectively. (b) Resonance fields of individual modes versus in-plane angle. (c) FMR spectra\ncalculated for individual single-layer discs with d0=150 nm and d\u00190:85d0, corresponding to bottom and top layers of SAF nanodiscs. Insets\nshow excitation regions for two eigen-modes. (d) Calculated FMR spectrum for individual SAF particle with d0andd\u00190:85d0. Insets display\nexcitation regions in both layers of SAF nanodisc. All spectra were calculated with in-plane field at jH=0\u000e(black arrow in left inset to c). (e)\nIn-plane angular dependence of resonance fields simulated for SAF disc array with d0=150 nm, d\u00190:85d0, and array periodicity 200 nm.\nthe temperature is lowered from room to below the Curie\npoint of the NiCu spacer. Our SAF design is rather unique\nas it effectively transforms the nano-elements from trilayers\nto single-layers when the spacer transitions from para- to fer-\nromagnetic state at about 220 K [29, 30, 38, 39]. The as-\nsociated transition in the FMR spectrum is very illuminat-\ning – the three-mode resonance is reduced to effectively two\nmodes, which is a known characteristic of single-layer nano-\narrays. Figure 3(b) shows the fits to the extracted high-order\nresonances (after subtracting the main peak): the pronounced\ndouble-resonance at room temperature with dominating L e1\n(red) and essentially a single broad peak at 170 K where the\nLe1peak is <1% of its RT value. The transition from two high-\norder peaks to one takes place precisely at the Curie temper-\nature of the spacer, deduced independently from our MOKE\nmagnetization data shown in Fig. 3(c). We thus confirm, in a\ndirect experiment, that the origin of the double spin-wave res-\nonance is unique to SAF nano-arrays. Specifically, the high-\norder (non-uniform) double resonance is a result of hybridiza-\ntion of predominantly acoustic and optical (in-phase and out-\nphase) oscillations in the two ferromagnetic layers comprising\nthe SAF. These modes collapse into one on direct exchange\ncoupling the outer layers below the TCof the spacer and, gen-\nerally, can be tuned by changing the SAF geometry and the\nspacer properties (via magnetic dilution and/or temperature).\nFIG. 3. (a) FMR spectra of SAF arrays measured at select tempera-\ntures. (b) Fits of high-order resonances (L e1,Le2) at above and below\nspacer’s Curie point. (c) Normalized remnant magnetization versus\ntemperature obtained from MOKE loops.\nThe above results demonstrate that the magnetization dy-\nnamics of periodic arrays of SAF nanomagnets have key dis-\ntinctions from the known behavior of single-layered nano-4\narrays. Our variable field-angle and temperature FMR and\nMOKE studies, supported by morphological characterization\nand micromagnetic simulations, enable us to identify the ori-\ngins of the distinct SAF properties observed. The three FMR\nmodes and their pronounced in-plane anisotropy are two such\ndistinct properties that, however, are governed by different\nfactors. Our micromagnetic simulations indicate that the split-\nting of the higher-order edge mode into two is due to the\ndipolar coupling within the individual, asymmetric SAF par-\nticles. Here, the inter-particle interaction within the array is\na secondary effect, which on the other hand dominates the\nanisotropic properties of the array. We find that the high-\norder spin excitations are hybridized acoustic (predominantly\nin-phase for L e1) and optical (predominantly out-of-phase for\nLe2) oscillations in the two layers comprising the SAF.\nThe occurrence of similar edge modes in single-layer nan-\nodiscs is usually explained by a non-uniform internal field dis-\ntribution, with its minima located at the edges orthogonal to\nthe static magnetization of the discs [14]. Following similar\nlogic for our asymmetric SAFs, we associate the two edge res-\nonances with non-uniform field distributions that are different\nin each of the two nanodiscs owing to their different internal\nand external (inter-layer) demagnetization field profiles. The\ndemagnetization is stronger in the smaller-diameter top layer,\nwhich results in a higher external field needed to excite the\nedge mode in this layer. In a symmetric SAF, with symmetric\ndemagnetization fields, only one edge mode is excited.\nThe four-fold anisotropy of the three FMR modes is due to\nthe bi-axial symmetry of our periodic arrays forming a square\nlattice. For our circular 150-nm SAF nanodisc arrays, this\nanisotropy is clearly caused by the inter-particle dipolar inter-actions. The observed 45\u000e-angle between the easy axes of the\nmain mode and the edge modes can be explained by the dif-\nference in the spacing of the respective excitation nodes for\nsingle-layer nanodot arrays [34]. We are able to recreate this\ntype of anisotropy in our micromagnetic simulations of SAF\nnano-arrays (to be discussed elsewhere).\nIn summary, we have investigated experimentally and mi-\ncromagnetically the spin dynamics of arrays of three-layer\nSAF nanoparticles where the intra-SAF coupling can be in-\nsitu controlled by varying temperature. The results show\nhow the intra- and inter-particle dipolar interactions com-\nbine to produce rather unique GHz properties of the system,\nsuch as additional spin-wave modes that can be tuned by the\nSAF geometry and/or temperature, which are relevant for the\nemerging field of nanostructured magnetic metamaterials and\nmagnonic devices.\nACKNOWLEDGMENTS\nWe thank Dr. Roman Verba for fruitful discussions. Sup-\nport from the Swedish Research Council (VR 2018-03526),\nthe Olle Engkvist Foundation (2020-207-0460), the V olkswa-\ngen Foundation (90418), and the National Academy of Sci-\nences of Ukraine (Projects 0119U100469, and 0120U100457)\nare gratefully acknowledged.\nDATA A VAILABILITY\nThe data supporting the findings of this study are available\nfrom the corresponding author upon request.\n[1] A. V . Chumak, V . I. Vasyuchka, A. A. Serga, and B. Hille-\nbrands, “Magnon spintronics,” Nature Physics 11, 453–461\n(2015).\n[2] X. M. Liu, J. Ding, G. N. Kakazei, and A. O. Adeyeye,\n“Magnonic crystals composed of Ni80Fe20 film on top of\nNi80Fe20 two-dimensional dot array,” Applied Physics Letters\n103, 062401 (2013).\n[3] B. Lenk, H. Ulrichs, F. Garbs, and M. Münzenberg, “The\nbuilding blocks of magnonics,” Physics Reports 507, 107–136\n(2011).\n[4] Arabinda Haldar and Adekunle Olusola Adeyeye, “Determinis-\ntic control of magnetization dynamics in reconfigurable nano-\nmagnetic networks for logic applications,” ACS Nano 10,\n1690–1698 (2016).\n[5] Susmita Saha, Ruma Mandal, Saswati Barman, Dheeraj Kumar,\nBivas Rana, Yasuhiro Fukuma, Satoshi Sugimoto, YoshiChika\nOtani, and Anjan Barman, “Tunable magnonic spectra in two-\ndimensional magnonic crystals with variable lattice symmetry,”\nAdvanced Functional Materials 23, 2378–2386 (2012).\n[6] Zhigang Liu, Richard D. Sydora, and Mark R. Freeman,\n“Shape effects on magnetization state transitions in individ-\nual 160-nm diameter Permalloy disks,” Physical Review B 77,\n174410 (2008).\n[7] M. L. Schneider, J. M. Shaw, A. B. Kos, Th. Gerrits, T. J.\nSilva, and R. D. McMichael, “Spin dynamics and damp-\ning in nanomagnets measured directly by frequency-resolvedmagneto-optic Kerr effect,” Journal of Applied Physics 102,\n103909 (2007).\n[8] R. V . Verba, E. G. Galkina, V . S. Tiberkevich, A. N. Slavin,\nand B. A. Ivanov, “Spin-wave modes localized on isolated de-\nfects in a two-dimensional array of dipolarly coupled magnetic\nnanodots,” Physical Review B 102, 054421 (2020).\n[9] C. Mathieu, C. Hartmann, M. Bauer, O. Buettner, S. Riedling,\nB. Roos, S. O. Demokritov, B. Hillebrands, B. Bartenlian,\nC. Chappert, D. Decanini, F. Rousseaux, E. Cambril, A. Müller,\nB. Hoffmann, and U. Hartmann, “Anisotropic magnetic cou-\npling of permalloy micron dots forming a square lattice,” Ap-\nplied Physics Letters 70, 2912–2914 (1997).\n[10] G. N. Kakazei, Yu. G. Pogorelov, M. D. Costa, T. Mewes, P. E.\nWigen, P. C. Hammel, V . O. Golub, T. Okuno, and V . Novosad,\n“Origin of fourfold anisotropy in square lattices of circular fer-\nromagnetic dots,” Physical Review B 74, 060406 (2006).\n[11] Giovanni Carlotti, “Pushing down the lateral dimension of sin-\ngle and coupled magnetic dots to the nanometric scale: Charac-\nteristics and evolution of the spin-wave eigenmodes,” Applied\nPhysics Reviews 6, 031304 (2019).\n[12] R. A. Duine, Kyung-Jin Lee, Stuart S. P. Parkin, and\nM. D. Stiles, “Synthetic antiferromagnetic spintronics,” Nature\nPhysics 14, 217–219 (2018).\n[13] Arezoo Etesamirad, Rodolfo Rodriguez, Joshua Bocanegra,\nRoman Verba, Jordan Katine, Ilya N. Krivorotov, Vasyl Ty-\nberkevych, Boris Ivanov, and Igor Barsukov, “Controlling5\nmagnon interaction by a nanoscale switch,” ACS Applied Ma-\nterials & Interfaces 13, 20288–20295 (2021).\n[14] Maurizio Pauselli, Andrzej A. Stankiewicz, and Giovanni Car-\nlotti, “Linear and non-linear dynamics of the free and reference\nlayers in a sub-40 nm magnetic tunnel junction: a micromag-\nnetic study,” Journal of Physics D: Applied Physics 50, 455007\n(2017).\n[15] A. Kamimaki, S. Iihama, K. Z. Suzuki, N. Yoshinaga, and\nS. Mizukami, “Parametric amplification of magnons in syn-\nthetic antiferromagnets,” Physical Review Applied 13, 044036\n(2020).\n[16] Jyotirmoy Chatterjee, Stephane Auffret, Ricardo Sousa, Paulo\nCoelho, Ioan-Lucian Prejbeanu, and Bernard Dieny, “Novel\nmultifunctional RKKY coupling layer for ultrathin perpendicu-\nlar synthetic antiferromagnet,” Scientific Reports 8, 1–9 (2018).\n[17] A. Talapatra and A. O. Adeyeye, “Coupled magnetic nanostruc-\ntures: Engineering lattice configurations,” Applied Physics Let-\nters118, 172404 (2021).\n[18] G. Carlotti, S. Tacchi, G. Gubbiotti, M. Madami, H. Dey,\nG. Csaba, and W. Porod, “Spin wave eigenmodes in single\nand coupled sub-150 nm rectangular permalloy dots,” Journal\nof Applied Physics 117, 17A316 (2015).\n[19] Arabinda Haldar and Adekunle Olusola Adeyeye, “Reconfig-\nurable and self-biased magnonic metamaterials,” Journal of Ap-\nplied Physics 128, 240902 (2020).\n[20] Krishna Begari and Arabinda Haldar, “Bias-free giant tunability\nof microwave properties in multilayer rhomboid nanomagnets,”\nJournal of Physics D: Applied Physics 51, 275004 (2018).\n[21] Jack C. Gartside, Alex Vanstone, Troy Dion, Kilian D. Sten-\nning, Daan M. Arroo, Hidekazu Kurebayashi, and Will R.\nBranford, “Reconfigurable magnonic mode-hybridisation and\nspectral control in a bicomponent artificial spin ice,” Nature\nCommunications 12, 1–9 (2021).\n[22] H. T. Nembach, R. D. McMichael, M. L. Schneider, J. M.\nShaw, and T. J. Silva, “Comparison of measured and simulated\nspin-wave mode spectra of magnetic nanostructures,” Applied\nPhysics Letters 118, 012408 (2021).\n[23] Brian B. Maranville, Robert D. McMichael, and David W.\nAbraham, “Variation of thin film edge magnetic properties with\npatterning process conditions in Ni80Fe20 stripes,” Applied\nPhysics Letters 90, 232504 (2007).\n[24] Han-Jong Chia, Feng Guo, L. M. Belova, and R. D.\nMcMichael, “Two-dimensional spectroscopic imaging of in-\ndividual ferromagnetic nanostripes,” Physical Review B 86,\n184406 (2012).\n[25] Zhizhi Zhang, Michael V ogel, M. Benjamin Jungfleisch, Axel\nHoffmann, Yan Nie, and Valentine Novosad, “Tuning edge-\nlocalized spin waves in magnetic microstripes by proximate\nmagnetic structures,” Physical Review B 100, 174434 (2019).\n[26] X. K. Hu, H. Dey, N. Liebing, H. W. Schumacher, G. Csaba,\nA. Orlov, G. H. Bernstein, and W. Porod, “Coherent precession\nin arrays of dipolar-coupled soft magnetic nanodots,” Journal of\nApplied Physics 117, 243905 (2015).\n[27] J. Ding, M. Kostylev, and A. O. Adeyeye, “Broadband fer-\nromagnetic resonance spectroscopy of permalloy triangularnanorings,” Applied Physics Letters 100, 062401 (2012).\n[28] Justin M. Shaw, T. J. Silva, Michael L. Schneider, and\nRobert D. McMichael, “Spin dynamics and mode structure in\nnanomagnet arrays: Effects of size and thickness on linewidth\nand damping,” Physical Review B 79, 184404 (2009).\n[29] A. F. Kravets, A. N. Timoshevskii, B. Z. Yanchitsky,\nM. A. Bergmann, J. Buhler, S. Andersson, and V . Ko-\nrenivski, “Temperature-controlled interlayer exchange coupling\nin strong/weak ferromagnetic multilayers: A thermomagnetic\nCurie switch,” Physical Review B 86, 214413 (2012).\n[30] A. F. Kravets, Yu. I. Dzhezherya, A. I. Tovstolytkin, I. M.\nKozak, A. Gryshchuk, Yu. O. Savina, V . A. Pashchenko, S. L.\nGnatchenko, B. Koop, and V . Korenivski, “Synthetic ferrimag-\nnets with thermomagnetic switching,” Physical Review B 90,\n104427 (2014).\n[31] Dmytro Polishchuk, Yuliya Tykhonenko-Polishchuk, Vla-\ndyslav Borynskyi, Anatolii Kravets, Alexandr Tovstolytkin,\nand Vladislav Korenivski, “Magnetic hysteresis in nanostruc-\ntures with thermally controlled RKKY coupling,” Nanoscale\nResearch Letters 13, 1–7 (2018).\n[32] B. C. Koop, T. Descamps, E. Holmgren, and V . Korenivski,\n“Relaxation-free and inertial switching in synthetic antiferro-\nmagnets subject to super-resonant excitation,” IEEE Transac-\ntions on Magnetics 53, 1–5 (2017).\n[33] Erik Holmgren, Artem Bondarenko, Bjorn Koop, Boris Ivanov,\nand Vladislav Korenivski, “Non-degeneracy and effects of pin-\nning in strongly coupled vortex pairs,” IEEE Transactions on\nMagnetics 53, 1–5 (2017).\n[34] G. N. Kakazei, X. M. Liu, J. Ding, V . O. Golub, O. Y . Salyuk,\nR. V . Verba, S. A. Bunyaev, and A. O. Adeyeye, “Large\nfour-fold magnetic anisotropy in two-dimensional modulated\nNi80Fe20 films,” Applied Physics Letters 107, 232402 (2015).\n[35] R. D. McMichael and M. D. Stiles, “Magnetic normal modes of\nnanoelements,” Journal of Applied Physics 97, 10J901 (2005).\n[36] Arne Vansteenkiste, Jonathan Leliaert, Mykola Dvornik, Math-\nias Helsen, Felipe Garcia-Sanchez, and Bartel Van Waeyen-\nberge, “The design and verification of MuMax3,” AIP Ad-\nvances 4, 107133 (2014).\n[37] Lukas Exl, Simon Bance, Franz Reichel, Thomas Schrefl,\nHans Peter Stimming, and Norbert J. Mauser, “LaBonte's\nmethod revisited: An effective steepest descent method for mi-\ncromagnetic energy minimization,” Journal of Applied Physics\n115, 17D118 (2014).\n[38] A. F. Kravets, A. I. Tovstolytkin, Yu I. Dzhezherya, D. M.\nPolishchuk, I. M. Kozak, and V . Korenivski, “Spin dynamics\nin a Curie-switch,” Journal of Physics: Condensed Matter 27,\n446003 (2015).\n[39] A. F. Kravets, D. M. Polishchuk, Yu. I. Dzhezherya, A. I. Tovs-\ntolytkin, V . O. Golub, and V . Korenivski, “Anisotropic mag-\nnetization relaxation in ferromagnetic multilayers with variable\ninterlayer exchange coupling,” Physical Review B 94, 064429\n(2016)." }, { "title": "1207.4522v1.Interface_Ferromagnetism_in_a_SrMnO3_LaMnO3_Superlattice.pdf", "content": "arXiv:1207.4522v1 [cond-mat.str-el] 18 Jul 2012INTERFACE FERROMAGNETISM IN A SrMnO 3/LaMnO 3SUPERLATTICE\nS. Smadici,1B.B. Nelson-Cheeseman,2A. Bhattacharya,2,3and P. Abbamonte1\n1Frederick Seitz Materials Research Laboratory, Universit y of Illinois, Urbana, Illinois 61801\n2Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439\n3Center for Nanoscale Materials, Argonne National Laborato ry, Argonne, Illinois 60439\nResonantsoft x-rayabsorption measurementsat theO Kedge onaSrMnO 3/LaMnO 3superlattice\nshow a shoulder at the energy of doped holes, which correspon ds to the main peak of resonant\nscattering from the modulation in the doped hole density. Sc attering line shape at the Mn L3,2\nedges has a strong variation below the ferromagnetic transi tion temperature. This variation has a\nperiod equal to half the superlattice superperiod and follo ws the development of the ferromagnetic\nmoment, pointing to a ferromagnetic phase developing at the interfaces. It occurs at the resonant\nenergies for Mn3+and Mn4+valences. A model for these observations is presented, whic h includes\na double-exchange two-site orbital and the variation with t emperature of the hopping frequency tij\nbetween the two sites.\nI. INTRODUCTION\nDoped La 1−xSrxMnO3(LSMO) has multiple FM,\nAFM and canted magnetic orders as a function of dop-\ningandtemperature1–3fromsuperexchange6anddouble-\nexchange4,5interactions, which favor an antiferromag-\nnetic (AFM) insulating phase and a ferromagnetic (FM)\nmetallic phase, respectively. The FM phase at low tem-\nperatures is near the x= 0.33 doping, at which the\nclosely related manganite La 1−xCaxMnO3shows very\nlarge (“colossal”) magnetoresistance (CMR)7.\nThe wave vectors of AFM and orbital orders in bulk\nmanganites can be accessed with soft x-ray scattering\nat the Mn L3,2edges. For instance, studies were made\nfor Pr 1−xCaxMnO3(Ref. 8), La 2−xSrxMnO4(Refs. 9–\n11)andLa 2−2xSr1+2xMn2O7(Ref.12)manganites. More\nrecent measurements showed that magnetic and orbital\nscattering amplitudes are similar13, studied the dop-\ning dependence14, confirmed the separation in energy\nof Ref. 13 between the magnetic and orbital scattering\nresonances,15and studied their evolution after photoex-\ncitation16. The broader features of the measurements\nare obtained in calculations of line shapes at the Mn\nL3,2edges with atomic multiplet models of magnetic17\nand orbital18scattering, and more recently, with a finite-\ndifference method19. However, the investigation of the\nbulk FM phase near the x= 0.33 doping is not possible\nwith soft x-ray scattering at the Mn Ledges, due to lack\nof contrast for this order.\nThe FM phase can be studied with soft x-ray scat-\ntering in (SMO)n/(LMO)2nsuperlattices (SL), in which\nthe Sr and La atoms are ordered in SrMnO 3(SMO) and\nLaMnO 3(LMO) layers. External magnetic fields are not\nnecessary. The SL growth sequence can be used to define\nthe period and symmetry of a reflection along the c-axis.\nThis was demonstrated for a n= 4 SL, in which the\nscattering wave vector was decreased to a range accessi-\nble at the O Kedge resonance and a higher symmetry of\na reflection used to probe interface scattering.20\nIn this work, we have applied these ideas to SL reflec-\ntions at the O Kand MnL3,2edges and studied thedevelopment of the FM moment in a shorter superpe-\nriodn= 2 SL with soft x-ray absorption and scatter-\ning. Measurements at the O Kedge showed modulated\nhole doping at oxygen sites. We have observed scatter-\ning at the Mn Ledges from the SL interfaces at SL re-\nflectionL= 2, following the temperature dependence\nof the FM moment. The symmetry of the SL reflec-\ntion allowed us to probe all Mn valences in the inter-\nface layers. In addition to Mn3+valence resonances of\nbulk AFM order measurements8–16, a peak in the reso-\nnant line shape, which has not been observed before, is\naligned with the fluorescence yield edge for the Mn4+va-\nlence. We present a model of the x-ray scattering from\nthe SL interfaces, which includes the temperature depen-\ndence of the double-exchange hopping frequency tijand\nthe change in the configurations of the Mn ions in the\nFM state.\nII. EXPERIMENTS\nA. Structure\nThen= 2 SL was grown on the (001) surface of\nSrTiO 3(STO)atArgonneNationalLaboratoryat700◦C\nin a 2×10−6Torr ozone pressure, followed by cool-\ning to 100◦C and pump down. The structure was\n{r×[2(SMO) + 4(LMO)] + SMO }withr= 13 [Fig.\n1(a)]. SMO ( aSMO= 3.805˚A, Ref. 21) and LMO\n(aLMO= 3.99˚A, Ref. 22) layers on the STO substrate\n(as= 3.905˚A) are under +2 .6 % tensile and −2.2 %\ncompressive strain. The surface RMS roughness, mea-\nsured with an atomic force microscope, was σs= 2.85˚A.\nFrom hard and soft x-ray reflectivity measurements, the\nSL superperiod was cSL= 22.5±0.5˚A and the average\nc-axisparameterfor1ML(acoverageofoneformulaunit\nof SMO or LMO over a as×bsarea) was 3 .86±0.05˚A.2\n(a) \n0.9 1.0 1.1 1.8 2.0 2.2 0510 15 Mn edge (643.6 eV) \n O edge (531.8 eV) \n 8048 eV Intensity (arb.units) \nL (r.l.u.) \n (c) I\nII L\nL\nS\nS\nL\nL\nL\n3+ \n3.5+ \n4+ \n3.5+ \n3+ \n3+ Sf\nifnif\nac\nnifif\nLf\n(b) L=1 \nL=2 \nπ−pol θ2θ b\naφbay\nx\nπσ\nFIG. 1: (Color online) (a) Sketch of a 6 ML superperiod of\nthe (SMO) 2/(LMO) 4superlattice with LaO ( L) and SrO ( S)\nplanes, layer form factors f, Mn valences in the MnO 2planes\nestimated from the neighboring LorSplanes, a magnetic\norder in the FM state and periods of the L= 1 andL= 2 SL\nreflections. (b) Scattering geometry for azimuth φ= 0◦. The\nscattering vector is along the c-axis, normal to the surface.\n(c) SL momentum scans at the Mn L3, OKedges and with\nhard x-rays (8048 eV).\nB. X-ray absorption\nX-ray absorption spectroscopy (XAS) measurements\nin fluorescence (FY) and electron yield (EY) modes were\nmade at undulator beamline X1B at the National Syn-\nchrotronLightSource. Theincidentlightwas π-polarized\nand the incidence and detector angles were θ= 80◦and\n2θ= 110◦[Fig. 1(b)]. The calculated energy resolution\nwas 0.39 eV and 0 .59 eV at O K(520 eV) and Mn L3\n(640 eV) edges, respectively.\nSL FY and EY measurements at the O Kedge show\ndoped holes on the oxygen sites (Fig. 2). Because the\nprobing depth exceeds the total SL thickness, FY has\ncontributions from both the SL and the substrate. In\ncontrast, because of the short electron escape depth, EY\nmeasurementsarefromtheSLtoplayersonly. Theshoul-der in FY measurements at 530.3 eV is aligned with the\nfirst peak in EY and is not present in FY measurements\nof the bare STO substrate. This shoulder corresponds to\ndoped holes in LSMO (Ref. 24) and to the L= 1 scatter-\ning peak at 529.6 eV. The peak at 531.8 eV is from the\nSTO substrate. The SL FYand EY measurementsat the\nOKedgeshownodiscerniblevariationwith temperature\nbetween 300 K and 255 K.\nSL FY and EYmeasurements at the Mn L3,2edges are\ncompared to FY measurements on bulk samples with dif-\nferent Mn valences (Refs. 25–27) in Fig. 3. SL FY does\nnot have the sharp peak at the lower energies character-\nistic of the Mn2+valence, which shows that the SL has\nonly the Mn3+and Mn4+valences. The SL FY measure-\nment was aligned in energy with the average of bulk FY\nfor the Mn3+and Mn4+valences, according to the num-\nber of SMO (Mn4+valence) and LMO (Mn3+valence)\nlayers in one superperiod. No discernible variation was\nobserved in FY or EY between 300 K and 255 K.\nXAS measurementsprobethe averagevalenceofOand\nMn atoms in the SL. To discern a variation with tem-\nperature in different SL layers it is necessary to turn to\nscattering.\nC. Resonant soft x-ray scattering\nResonant soft x-ray scattering (RSXS) measurements\nwere made at the same beamline with an ultra-high\nvacuum diffractometer. For other RSXS experiments\non bulk and SL at this endstation see Refs. 20,23,28–\n31. Detection was for both π−πandπ−σchannels.\nThe scattering momentum in the reflectivity geometry\nQ= (0,0,2πL/cSL) is indexed with respect to the SL\nsuperperiod cSL. The energy resolution was 0 .20 eV and\n0.34 eV at O K(520 eV) and Mn L3(640 eV) edges, re-\nspectively. The sample was cooled in zero magnetic field\nand scattering measurements for L= 1 at the O Kedge\nand forL= 1,2 at the Mn L3,2edges [Fig. 1(c)] were\nmade at different temperatures (Figs. 2 and 4).\nFM order in metallic films has been studied with x-\nray resonant magnetic scattering of linearly32and cir-\ncularly33polarized light at the Fe and Co Ledges. The\nFMorderinaAg/NiSLhasbeeninvestigatedwithcircu-\nlarly polarized light34. However, unlike previous studies,\nwhere an external magnetic field was applied to separate\nthe magnetic and charge scattering, the SL FM order is\naccessed here at SL reflections with no applied magnetic\nfields.\n1. OKedge line shape\nThe spatial modulation in the density of holes doped\non the oxygen sites can be observed with RSXS. Specifi-\ncally, the O Kedge line shape for L= 1 scattering shows\na peak closeto the energyof the shoulderin FY measure-\nments (Fig. 2).3\n525 530 535 540 545 0510 15 020 40 \n330 K 310 K 290 K 270 K 255 K 185 K 160 K 135 K 110 K 90 K RSXS intensity \n(arb. units) \nEnergy (eV) FY and EY \n(arb. units) FY (SL+STO at 300 K) \n FY (SL+STO at 255 K) \n FY (STO only) \n EY (SL only) \n \n \n \nSTO \nSL \neL,S \ng,\neL\ng,eL,S \ng,\n2 ,gt↑,ge↑2 ,gt↓,ge↓\n2-3 eV 1-2 eV \nFIG. 2: (Color online) Top: FY at the O Kedge for the\nSL on STO at different temperatures and the bare STO\nsubstrate (black line), compared to EY measurements. FY\nmeasurements were aligned with the linear transformation\nFYplotted=aFYmeasured +b, whereaandbare constants.\nBottom: Temperature dependence of resonant scattering at\nL= 1. The scans have been normalized tothe pre-edge values\nand shifted vertically for clarity. The inset shows the orde r\nof the energy levels for SMO in the ground state, with the\nJahn-Teller splitting neglected. The upper level eg,↓is split\nin LMO.\nSince the order of levels in the RSXS line shape at the\nOKedge follows that of the ground state, the RSXS line\nshape canbe analyzedusingthe hybridizationbetween O\npand Mneglevels in the ground state. XAS experiments\nand calculationsat the Mn Ledge(Ref. 35) givea crystal\nfield splitting 10 Dqbetween the egandt2glevels of 1.5\neV for bulk LMO and 2.4 eV for SMO. Fig. 2 (inset)\nshows the unoccupied eg,↑,t2g,↓andeg,↓levels in SMO.\nThe scattering line shape at the O edge is described well\nby Opstates hybrids with the Mn eglevels in the SMO\nand LMO layers, shown with arrows in Fig. 2, followed\nat higher energy by hybrids with La and Sr states. Two\neg,↓levels (3.5 eV and 5 eV above eg,↑) are present in\nLMO and only one for SMO (3.5 eV above eg,↑) (Ref. 36)\nbecause the electron in the eg,↑level in LMO splits the\nunoccupied eg,↓levels by Coulomb interaction, even in\nthe absence of any Jahn-Teller distortion.\nHowever, no variation across the FM transition is ob-\nserved within the error bars. To access T-dependent in-635 640 645 650 655 0246FE D\nEY 205 K EY 300 K Mn 2+ \n Mn 3+ \n Mn 4+ L=2 (T=90 K) CBA\n Intensity (arb. units) \nEnergy (eV) FY \n(T=300 K )\n FY \n(T=255 K )\nFIG. 3: (Color online) SL EY and FY (lower curves) and\nRSXS (top) at L= 2 at the Mn L3,2edges compared to\nFY measurements for different Mn valences (middle, from\nRef. 26).\nterface states (Sec. III A), it is necessary to reach the\nL= 2 reflection, for which the O Kedge energy is too\nlow. In contrast, L= 2isaccessibleatthe Mn L3,2edges.\n2. MnL3,2edge valences\nThere is no discernible variation in the line shape at\nL= 1 at the Mn L3,2edges across the FM transition\ntemperature (Fig. 4). However, a pronounced variation\nis visible for L= 2. From under the relatively broad\nXAS at the Mn Ledge, T-dependent RSXS at L= 2\nselects those states that are sensitive to the temperature\nvariation. Specifically, an increased intensity of the A,\nBandCpeaks at lower T, and a decreased intensity of\ntheαpeak is observed at the L3edge (Fig. 4). Parallel\nvariations occur at the L2edge for peaks E,Fandβ.\nSimilar results were obtained for azimuthal angle φ=\n45◦(Fig. 5), consistent with the azimuthal dependence\nof magnetic scattering. The temperature dependence of\nheight and width of peak Care shown in Fig. 6.\nCoulomb and exchange interactions for ground and\nRSXS excited states are different at the Mn L3,2edges.\nAn analysis of the RSXS line shapes based on ground\nstatecalculations,similartothatattheO Kedge,cannot\nbe made. However, both FY and RSXS measurements4\nprobe excited states and FY measurements on bulk sam-\nples for different Mn valences will be used to identify the\nvalence of scatterers for different resonances in the RSXS\nline shapes. This approach is supported by the relatively\nsmall difference in energy between RSXS resonances for\nscatterers of the same valence and different specific scat-\ntering contrasts ( /lessorsimilar1 eV between magnetic and orbital\nscattering for bulk measurements13,15) compared to the\n∼2 eV separation in energy between the main FY fea-\ntures for different valences (Fig. 3).\nFY measurements in Sec. II B showed that the Mn2+\nvalence is absent in the SL. The AandBresonances\nin the RSXS line shape for L= 2 correspond to the\nMn3+valence, while resonance Clines up at the energy\nof FY edge for a Mn4+valence. More T-dependent mea-\nsurements on SL with different superperiods are needed\nbefore a quantitative discussion of peak α.\nIII. DISCUSSION\nThe model of Sec. III A relates the variation with\ntemperature of the scattering contrast for L= 2 and the\nabsence of variation for L= 1 to changes in the form fac-\ntorδfiof interface layers and interface ferromagnetism.\nThe variation of δfi=fi,FM−fi,PMwith the transi-\ntion from the PM to the FM state for specific interface\nmagnetic and orbital x-rayscattering models is discussed\nin Sec. III B.\nA. Interface ferromagnetism\nThe line shape and structure factor at the Mn edge at\nL= 2(Fig. 4)aremadeoftwoparts. TheT-independent\nstructurefactor Sis givenbythe SLstructure, to whicha\nT-dependentcontribution δSisaddedwiththetransition\nto the FM state [Fig. 6(a)].\nThe scattering intensity follows the evolution with\ntemperature of the FM moment [Fig. 6(a)]. The pos-\nsibility that the transition to the FM state gives exclu-\nsively non-magnetic scattering contrast is not supported\nby an analysis of the line shapes (Sec. III B). The T-\ndependent scattering δSis, at least partially, magnetic in\norigin and, for simplicity, we discuss only x-ray resonant\nmagnetic scatteringin this section. Orbitalcontributions\nto temperature variations in line shapes are addressed in\nSec. III B.\nThe scattering intensity is I∝/vextendsingle/vextendsingleS+δS/vextendsingle/vextendsingle2. The scat-\ntering form factors ( f)mnand (δf)mnare tensors, which\nare multiplied with the final (ˆ ǫf) and initial (ˆ ǫi) light po-\nlarization vectors. This gives an overall factor which, for\ncharge (f) and magnetic ( δf) scattering, is\nS∝ˆǫ∗\nf,m(f)mnˆǫi,n∝(ˆǫ∗\nfˆǫi)f(ω) (1)\nδS∝ˆǫ∗\nf,m(δf)mnˆǫi,n∝i[(ˆǫ∗\nf׈ǫi)ˆzl]δf(ω) (2)635 640 645 650 655 660 0510 15 20 0510 \n330 K 310 K β\nL=1 \nL=2 α Intensity (arb. units) \n295 K 275 K 260 K 240 K 220 K 205 K 185 K 165 K 145 K 125 K 105 K 90 K \nEY 300 K \nEnergy (eV) \n 330 K 310 K 280 K 255 K 90 K \n Mn 3+ Mn 4+ \nFIG. 4: (Color online) Temperature dependence of RSXS at\nthe MnL3,2edges atL= 1 andL= 2 for azimuthal angle\nφ= 0◦.\nwhere ˆzlis the direction of the local moment at site l\n(Refs. 32,34). The sum over the in-plane sites lfor each\nlayeris proportionalto the magnetization− →Mofthe layer.\nf(ω) andδf(ω) are scalar functions.\nRSXS peaks that persist above the FM transition tem-\nperature are due to scattering contrast defined by the SL\nstructure, which is held constant by the internal field\nbetween Sr2+and La3+ions arranged in the SL layers.\nWith the form factors shown Fig. 1(a) for “interface”\n(fi), “near-interface” ( fni), “middle SMO” ( fS) and\n“middle LMO” ( fL) layers, and neglecting inter-diffusion\nroughnessand structuraldifferences between SMO/LMO\nand LMO/SMO interfaces, the T-independent structure\nfactorsS(Q) =/summationtext\nlfleiQzlatL= 1,2 are\nS(L= 1) =−fL+fS+fi−fni (3)\nS(L= 2) =fL+fS−fi−fni (4)\nThe origin has been chosen so that an arbitrary phase\nfactor between S(L= 1) andS(L= 2) is zero.\nHigher momenta Lare not accessible at the Mn L5\nedges. However, the absence of a variation with temper-\nature in the scatteringintensity at L= 1 [δS(L= 1) = 0]\nand the variation at L= 2 [δS(L= 2)∝ne}ationslash= 0] strongly sug-\ngests that the unit cell of the T-dependent contribution\nto the structure factor ( δS) is half the SL superperiod or\n3 ML. The middle of the SMO and LMO layers (sepa-\nrated by 3 ML) are the most dissimilar parts of the SL\nstructure, while the SMO/LMO and LMO/SMO inter-\nfaces (with two interfaces every superperiod, also sepa-\nrated by 3 ML) are similar. In the following, we consider\na scattering component δSthat develops at the SL inter-\nfaces.\nThese conditions are contrary to those expected for\nscattering from crystal field effects or structural differ-\nences in a SL, from either differences in the c-axis lattice\nconstant or Jahn-Teller distortions, which do not have a\n3 ML unit cell. In addition, no discernible variation was\nobserved for scattering at the La M5,4edges forL= 1\norL= 2 between 300 K and 225 K (data not shown).\nThis shows that the change in the line shape is due to a\nvariation with temperature in the resonant form factors\nof the SL layers ( δfl), not of structural factors ( δzl).\nIn addition, these symmetry conditions on δSset\nmore stringent constraints on the variation of δflin the\nSL, beyond the experimental observation δS(L= 1) =\n−δfL+δfS+δfi−δfni= 0. To obtain a 3 ML unit cell,\nthe variation in the middle of the LMO and SMO layers\nmust be equal, δfL=δfS. In addition, the variation\nin the interface and near-interface layers must be equal,\nδfi=δfni[Fig. 1(a)].\nFM order in the iandnilayers is consistent with esti-\nmates of the average Mn valence in a MnO 2plane based\non the type ( LorS) of neighboring planes [Fig. 1(a)].\nSpecifically, a comparison to magnetic orders of equiv-\nalent bulk LSMO doping shows that Mn3+and Mn3.5+\nvalencesareneartheFMdomeforbulkLSMO.Themag-\nnitude of the FM moment depends on the Mn valence\nand implicitly on the SL interface roughness, with struc-\ntural imperfections in a (SMO) 4.4/(LMO) 11.8SL corre-\nlatingwiththeaverageinterfaceFMmoment.38However,\nthe FM moment distribution is more symmetrical in the\nsmaller superperiod n= 3 SL (Ref. 39), consistent with\nthe symmetric FM moment distribution in Fig. 1(a).\nThe 3 ML unit cell of T-dependent scattering shows\nthat there must be two regions within a superperiod\nwhich are different from the FM iandnilayers. With\none the middle of the SMO layers ( fS), the other must\nbe the middle of the LMO layers( fL). The x-ray scatter-\ning measurements imply that the magnetic scattering in\nthesetwolayersis thesame( δfS)mag= (δfL)mag. There-\nfore, these layers have either the same magnetization− →M\nor no magnetization at low temperatures. The different\nhole doping of these layers does not support the possibil-\nity of an equal magnetization. The remaining possibility\nis that, as the SL is cooled and becomes FM in zero ap-\nplied field, there is no magnetization in both these layers,\nor (δfS)mag= (δfL)mag= 0. Therefore, for the n= 2\nSL, the variation in magnetic scattering ( δf)magand FM(a) \n(b) Imin Imax \nT=255 K \nT=300 K ABCEF\nD\nαβ\nFIG. 5: (Color online) Two-dimensional resonance profiles a t\nthe MnL3,2edges at 255 K and 300 K for azimuth φ= 45◦.\nphase are localized at the SL interfaces.\nA model of the magnetic state for the FM SL is shown\nin Fig. 1(a), where IandIIrepresent magnetic phases\nof thefSandfLlayers in no applied field (in con-\ntrast, the polarized neutron reflectivity measurements in\nRefs. 38,41 were made in applied fields). Since the aver-\nage magnetization is zero for both IandIIphases, the\nmagnetic scattering has a 3 ML period. There are sev-\neral different possible IandIIphases: a PM phase, an\nordered AFM phase (for instance, a C-type or a G-type),\nor an irregular phase with canted moments41pointing in\ndifferent directions in the sample regions with slight vari-\nations in local doping3, even though the moments in the\nfiandfnilayers are always parallel.\nThe valences of the fLandfSlayers are close to\nMn3+and Mn4+, which correspond to A-type and G-\ntype AFM magnetic orders in bulk LMO and SMO, with\ntransition temperatures of T N,LMO= 135 K (Ref. 37)\nand T N,SMO= 235 K (Ref. 21), respectively. This sug-\ngests PMIorIIphases, at least for the higher tem-\nperature range, below the SL FM transition tempera-\nture of 305 K [Fig. 6(a)]. However, the SL saturation6\nFM moment of ∼2.5µBat 5 K (Ref. 40) gives the ex-\ntent along the c-axis of the FM region in high fields of\n/greaterorsimilar(2.5/3.22)×6 ML∼4.65 ML for each superperiod,\nwhere∼3.22µBis the maximum FM moment of the\nx= 0.33 alloy39. This value is too high for both IandII\nphases to remain PM at the lowest temperatures. There-\nfore, at least one magnetic transition, other than the FM\ntransition, occurs in the SL.\nThe RSXS measurements are consistent with transi-\ntions in the fSandfLlayers from a PM phase at higher\ntemperatures to either a G-type or C-type AFM phase\n(near the Mn4+doping of bulk LSMO) and to an irreg-\nular canted phase (near the the Mn3+doping3) at lower\ntemperatures, respectively. However, since there is no\naverage layer magnetization in all these cases, the RSXS\nintensity does not vary at these transitions, in contrast\nto the FM transition in the iandnilayers.\nWith these constraints on δfl, the change in the struc-\nture factor δS(Q) =/summationtext\nlδfleiQzlat the FM transition\nandL= 1,2 from Eqs. (3)-(4) is\nδS(L= 1) = 0 (5)\nδS(L= 2) =−2δfi (6)\nEq. (6) relates the changes with temperature in the line\nshape atL= 2 to variations of form factor of interface\nδfiand near-interface δfni=δfilayers in the SL.\nTheδS(L= 2) reflectionis allowedin this n= 2 SL for\nall Mn sites in the FM layers. However, similar to AFM\norders in bulk LSMO, scattering from a Mn4+valence\nwasnotobservedat L= 3foran= 4SL (Ref. 20) [it was\nobserved in a n= 3 SL atL= 3 (data not shown)]. The\nsymmetry that very effectively forbids reflections from\nthe Mn4+ions atL= 3 for the n= 4 SL is not known\nandsurprising,giveninherentsmallimperfectionsofaSL\nstructure. More measurements are needed for different\nSL to answer this question.\nB. Interface x-ray scattering\nWe now discussthe temperaturevariationofthe RSXS\nline shape at L= 2 and Mn L3,2edges.\nThe width of resonance C, corresponding to the Mn4+\nvalence and to interface and near-interface layers, has an\nsharp increase at the FM transition temperature [Fig.\n6(b)]. The increase in the scattering intensity in the FM\nstate is also taking place ∼0.2 eV below the chargeorder\nresonance that corresponds to the Mn4+valence in the\nPM state (at 644 .65 eV in Fig. 4).\nIn general, the line shape of resonantmagnetic scatter-\ning is related to variations in the occupation of orbitals\ninduced by a magnetic field42near the FY edges for Mn\nions of different (Mn3+and Mn4+) valences. However,\nthe magnetic scattering is slightly shifted to lower en-\nergies compared to orbital scattering for AFM bulk or-\nders.13,15We cannot resolve two peaks at Cin the SL\nline shapes at low T, but this suggests that, with the(a) \n0 100 200 300 400 10 -4 10 -3 \n0.00 0.75 1.50 2.25 \n R(B=0 T) \n R(B=9 T) \n R (ohm-cm) \nT (K) \n Moment ( µB/f.u.) \n FC (200 Oe) \n ZFC \n0.00 0.25 0.50 0.75 \n Mn 4+ scattering \nheight (arb. units) \n \n(b) \n50 100 150 200 250 300 350 0.7 0.8 0.9 1.0 1.1 1.2 Mn 4+ scattering FWHM (eV) \nT (K) 1ψ\n2ψ\n()()Mn O Mn i j− − \nFIG. 6: (Color online) (a) Temperature dependence of res-\nonanceCheight forφ= 0◦compared to the SL FM mo-\nment measured with SQUID for ZFC and in-plane FC=200\nOe (Ref. 40). The SL has a 305 ±5 K FM transition tempera-\nture, which is lower than the ∼355 K transition temperature\nof thex= 0.33 LSMO alloy (Ref. 39). Hysteresis loops show\nthat the easy axis is in-plane (data not shown). SL resistanc e\nbecomes metallic-like at low T (Ref. 40). (b) Temperature\ndependence of resonance Cwidth. The line is a guide to the\neye. Inset shows a sketch of the double-exchange configura-\ntions|ψ1/angbracketrightand|ψ2/angbracketrightfor Mn sites iandj(Ref. 4).\nincrease of the FM moment at lower T, a T-dependent\nmagneticscatteringcontributionisadded ∼0.2eV below\nthe charge scattering resonance. This addition to fiof\na temperature dependent ( δfi)magexplains the observed\nvariation in line shape at L= 2. The charge scatter-\ning resonance might also increase at lower T, concomi-\ntantly with magnetic scattering and variations in orbital\nscattering with T are discussed briefly at the end of this\nSection.\nA more gradual increase in width is observed at lower\nT [Fig. 6(b)]. For x-ray scattering in the FM phase, it is\nnecessary to consider a double-exchange two-site orbital,\nwhichsuggeststhat thiswidth increaseisrelatedtotheT\ndependenceofthedouble-exchangefrequency tijbetween\nthe two Mn sites. Both resonant magnetic and orbital\nscattering are ultimately scattering off orbitals, and the\nconsideration of two-site orbitals in the FM state applies\nto both cases.\nThe double-exchange process involves two coordinated\njumps from the Mn to the O atoms [Fig. 6(b), inset].7\nIt is useful to consider the simpler process of one jump\nfirst, which is sometimes included in XAS calculations of\ncomplex oxides. In this case, inter-site charge transfer\nbetweend-states of a transition metal and a neighboring\n(ligand,L) O ion43and consideration of multiple con-\nfigurations (for instance, d8andd9Lfor Cu1+and Cu2+\nvalences) change the scattering form factor fat the tran-\nsition metal edge. In particular, satellite peaks develop\nin XAS (and, implicitly, in RSXS) at additional Cu va-\nlences.44\nIn the double-exchange process, specific to FM com-\nplex oxides, charge transfer takes place between transi-\ntionmetalsites,beyondtheneighboringOatoms. Specif-\nically, the |ψ1∝an}bracketri}htand|ψ2∝an}bracketri}htconfigurations are coupled in a\ntwo-site ground state wave function [Fig. 6(b), inset],\nwhich for this FM manganite is\n|ψ±∝an}bracketri}ht=1√\n2(|ψ1∝an}bracketri}ht±|ψ2∝an}bracketri}ht) (7)\n|ψ1∝an}bracketri}ht=|Mn3+,O2−,Mn4+∝an}bracketri}ht\n|ψ2∝an}bracketri}ht=|Mn4+,O2−,Mn3+∝an}bracketri}ht\nwith Mn valences in FM layers in a superposition of\nMn3+and Mn4+. In the ground state (without an x-\nray photon absorbed), the charge transfer splits the two\nlevels|ψ±∝an}bracketri}htby the exchange energy 2 tij(Ref. 4), where\nthe double-exchange hopping between sites iandjis\ntij=tDEcos[(θi−θj)/2], withtDEa constant and θi,j\nthet2g,↑spins orientations on the two sites (Ref. 5).\nTo account for the double-exchange process in x-ray\nscattering, the orbitals |ψ1,2∝an}bracketri}htare replaced with the two-\nsite orbitals |ψ±∝an}bracketri}ht. Similar to the case of ligand holes\non oxygen atoms, the charge transfer between Mn sites\nbeyond the neighboring O atoms changes the scattering\nfactorfat satellite peaks in RSXS, which correspond to\nthe Mn3+and Mn4+valences.\nIn addition, the splitting by 2 tijof the|ψ±∝an}bracketri}htstates\nresults, for our relatively large energy resolution and core\nhole width, in an increase of the measured width by 2 tij.\nMoreprecisely,thethebandwidthof egelectronsdepends\non the hopping frequency between the iandjsites as37\nW∝cos[(θi−θj)/2]cosφ∝tijcosφ (8)\nwhere (π−φ) is the angle between the Mn-O-Mn bonds.\nA width increase at lower T in the ground state is trans-\nferred to a increase of the RSXS line width.\nThe hoppingfrequency tijincreaseswith increasedFM\norder of spins θi,jat lower T, and broadens the scatter-\ning form factors fand the line width. In this model,\nthe XAS and therefore, the RSXS peaks, should become\nbroader at lower temperatures. The width increase at\nlower temperatures of peak C[Fig. 6(b)] is consistent\nwith this model and tband∼0.2−0.5 eV for each of the\negstates, 2tDE∼2TCurie∼0.05 eV and contributions\nfrom experimental resolution (0 .34 eV at the Mn Ledge)\nand core-hole width ( wFWHM∼0.3−0.5 eV, Ref. 8).\nInadditiontothedouble-exchangeprocessesintheFM\nstate, lattice distortions are also relevant to the CMRtransition45,46. In bulk manganites, they may depend on\nT,changingthebondalignment φandbandwidth W[Eq.\n(8)]. However, the average angle between the Mn-O-Mn\nbonds for SL samples is fixed by the substrate.\nOrbital scattering at the Mn Ledges has a compa-\nrable amplitude to magnetic scattering for bulk AFM\norders.13,15It can come from occupation contrast or po-\nlarization contrast from different atomic orbital orien-\ntations in the anomalous scattering tensor. The anal-\nogous occupation contrast in SL FM is a T-dependent\ncharge transfer across SL interfaces (which includes the\nelectronic reconstruction of Ref. 20), in addition to the\nT-independent part defined by the SL structure. The T-\ndependent polarization contrast in the SL may also be\nsubstantial; for instance, on closely related SL (Ref. 47),\nin-planeeg(x2−y2) occupation and FM near LMO in-\nterfaces and out-of-plane eg(3z2−r2) orbital occupation\nand AFM in the middle of LMO layers was inferred from\nXMLD and XMCD measurements. Polarization-resolved\nscattering measurements in a magnetic field with πand\nσincident light and scattered beam polarization analysis\nare necessary to separate different magnetic and orbital\ncontributions to scattering at the Mn L3,2edges.\nWe discuss the O Kedge briefly. Oxygen doping is\nconsistent with our observations (Fig. 2), other measure-\nments24and certain models14. The interface FM state of\nthisn= 2 SL is the metallic state observed in a n= 4\nSL (Ref. 20). The reflection at L= 1 (Fig. 2) is not sen-\nsitive to T-dependent scattering because, as for the Mn\nLedges,δS(L= 1) = 0. The intensity of the L= 3 re-\nflection in Ref. 20 corresponds to the L= 2 reflection for\nthis SL. Inparticular, to determinewhether T-dependent\nscattering occurs at the O Kedge in this SL, it would\nbe necessary to measure at L= 2, a momentum which is\nnot accessible at the O Kedge.\nIV. CONCLUSION\nX-ray absorption measurements at the O Kedge in a\nSrMnO 3/LaMnO 3superlattice showed a shoulder, corre-\nsponding to holes doped on oxygen sites. The shoulder is\nalignedwith the main resonantpeak ofsoft x-rayscatter-\ning from the spatial modulation in the density of doped\nholes.\nA large variation in the Mn L3,2line shapes at L= 2,\nbut not atL= 1, was observed across the FM transition,\npointing to scattering from ferromagnetic interfaces.\nComparison to fluorescence yield edge energies for dif-\nferentMn valencesshowedthe presenceofscatteringcon-\ntrast at both Mn3+and Mn4+valences. An x-ray scat-\ntering model, which includes double-exchange orbitals in\nthe FM state, explains the observed line broadening at\nlower temperatures.\nMeasurements on bulk 113 (Ref. 8), 214 (Ref. 9–11)\nand 327 (Ref. 12) manganites in the AFM state observe\ntwo main resonances only at the Mn L3edge, atAand\nB. Scattering from Mn4+ions (corresponding to res-8\nonanceC) has the symmetry of a forbidden reflection.\nSpecifically, it has a spatial periodicity of 2 u.c. and is\nnot allowed at the bulk in-plane orbital order reflection\nwavevector of 4 u.c. along the tetragonal axes (Ref. 8,9).\nHaving to rely on measurements of the Mn3+resonances\nonly, different methods to determine the charge dispro-\nportionation for bulk AFM orders are controversial, with\nboth small and large chargedisproportionation obtained.\nOur RSXS line shapes, for a SL structure with a large\nintrinsic charge disproportionation, add an experimental\nconstraint on these competing models.\nThe development of the SL FM order was accessed\nwith x-ray resonant magnetic scattering and no applied\nmagnetic fields. An open question is the trace [FC or\nZFC in Fig. 6(a)] that the height of resonance Cwould\nfollow on further cooling.\nWe would like to contrast our measurements to polar-\nized neutron reflectivity (PNR) data on SMO/LMO su-\nperlattices (Refs. 38,41), where a magnetic modulation\nwas measured with a period equal to the SL superpe-\nriod (magnetization strongly suppressed in SMO, high in\nLMO). In contrast, the RSXS measurements presented\nhereshowanorderingofmagneticmomentswithaperiod\nequal to half the SL superperiod. Several factors may be\nat the origin of this difference. First, the experimentalconditions of the PNR and RSXS measurements were\ndifferent. Specifically, PNR measurements were made in\nrelativelyhighfields (0 .55T and0.82T inRef. 38and41,\nrespectively), while the RSXS measurements were made\nwith no applied fields. Second, the samples measured in\nthis work have a lower SL superperiod ( n= 2) compared\nto the samples of PNR measurements ( n= 3 andn= 5).\nThus, a complete mapping of the magnetic structure of\nSMO/LMO superlattices as a function of deposition se-\nquence, magnetic field and temperature requires more\nmeasurements.\nV. ACKNOWLEDGMENTS\nThis work was supported by the Department of En-\nergy Office of Basic Energy Science: RSXS measure-\nments by grant DE-FG02-06ER46285,NSLS facilities by\nDE-AC02-98CH10886, and MRL facilities by DE-FG02-\n07ER46453andDE-FG02-07ER46471. WorkatArgonne\nNational Laboratory, including use of facilities at the\nCenter for Nanoscale Materials, was supported by the\nU.S. Department of Energy, Office of Basic Energy Sci-\nences under contract No. DE-AC02-06CH11357.\n1E.O. Wollan and W.C. Koehler, Phys. Rev. 100, 545\n(1955).\n2J.B. Goodenough, Phys. Rev. 100, 564 (1955).\n3P.-G. de Gennes, Phys. Rev. 118, 141 (1960).\n4C. Zener, Phys. Rev. 82, 403 (1951).\n5P.W. Anderson and H. Hasegawa, Phys. Rev. 100, 675\n(1955).\n6P.W. Anderson, Phys. Rev. 79, 350 (1950).\n7S. Jin, T.H. Tiefel, M. McCormack, R.A. Fastnacht, R.\nRamesh, and L.H. Chen, Science 264, 413 (1994).\n8K.J. Thomas et al., Phys. Rev. Lett. 92, 237204 (2004).\n9S.B. Wilkins et al., Phys. Rev. Lett. 91, 167205 (2003).\n10S.S. Dhesi et al., Phys. Rev. Lett. 92, 056403 (2004).\n11S.B. Wilkins et al., Phys. Rev. B 71, 245102 (2005).\n12S.B. Wilkins et al., J. Phys.: Condens. Matter 18, L323\n(2006).\n13U. Staub et al., Phys. Rev. B 79, 224419 (2009).\n14M. Garcia-Fernandez et al., Phys. Rev. B 82, 235108\n(2010).\n15S.Y. Zhou et al., Phys. Rev. Lett. 106, 186404 (2011).\n16H. Ehrke et al., Phys. Rev. Lett. 106, 217401 (2011).\n17N. Stojic, N. Binggeli, and M. Altarelli, Phys. Rev. B 72,\n104108 (2005).\n18C.W.M. Castleton and M. Altarelli, Phys. Rev. B 62, 1033\n(2000).\n19O. Bunau, PhD thesis, University of Grenoble (2010).\n20S. Smadici et al., Phys. Rev. Lett. 99, 196404 (2007).\n21O. Chmaissem et al., Phys. Rev. B 64, 134412 (2001).\n22J. Rodriguez-Carvajal et al., Phys. Rev. B 57, R3189\n(1998).\n23S. Smadici, J.C.T. Lee, J. Morales, G. Logvenov, O. Pelleg,\nI. Bozovic, Y. Zhu and P. Abbamonte, Phys. Rev. B 84,155411 (2011).\n24H.L. Ju, H-C. Sohn, and K.M. Krishnan, Phys. Rev. Lett.\n79, 3230 (1997).\n25S. P. Cramer et al., J. Am. Chem. Soc. 113, 7937 (1991).\n26F. Morales et al., J. Phys. Chem. B 108, 16201 (2004).\n27J. Leeet al., Phys. Rev. B 80, 205112 (2009).\n28P. Abbamonte, A. Rusydi, S. Smadici, G.D. Gu, G.A.\nSawatzky, and D.L. Feng, Nature Physics 1, 155-158\n(2005).\n29J.-S. Lee, C.-C. Kao, C.S. Nelson, H. Jang, K.-T. Ko, S.B.\nKim, Y.J. Choi, S.-W. Cheong, S. Smadici, P. Abbamonte,\nand J.-H. Park, Phys. Rev. Lett. 107, 037206 (2011).\n30S. Smadici, J.C.T. Lee, S. Wang, P. Abbamonte, G.\nLogvenov, A. Gozar, C. Deville Cavellin, and I. Bozovic,\nPhys. Rev. Lett. 102, 107004 (2009).\n31S. Smadici, J.C.T. Lee, A. Rusydi, G. Logvenov, I. Bo-\nzovic, and P. Abbamonte, Phys. Rev. B 85, 094519 (2012).\n32C.C. Kao et al., Phys. Rev. Lett. 65, 373 (1990).\n33C.C. Kao et al., Phys. Rev. B 50, 9599 (1994).\n34J.M. Tonnerre et al., Phys. Rev. Lett. 75, 740 (1995).\n35M. Abbate et al., Phys. Rev. B 46, 4511 (1992).\n36T. Saitoh et al., Phys. Rev. B 51, 13942 (1995).\n37J. Topfer et al., J. of Solid State Chemistry 130, 117\n(1997).\n38S. Mayet al., Phys. Rev. B 77, 174409 (2008).\n39A. Bhattacharya et al., Phys. Rev. Lett. 100, 257203\n(2008).\n40B.B. Nelson-Cheeseman et al.(unpublished).\n41T. Santos et al., Phys. Rev. Lett. 107, 167202 (2011).\n42J.P. Hannon, G.T. Trammell, M. Blume, and D. Gibbs,\nPhys. Rev. Lett. 61, 1245 (1988).\n43F. de Groot, Coordination Chemistry Reviews 249, 319\n(2005).\n44Z. Huet al., Chemical Physics 232, 63 (1998).\n45H.Y.Hwang, S.W.Cheong, P.G.Radaelli, M. Marezio, and\nB. Batlogg, Phys. Rev. Lett. 75, 914 (1995).46A.J. Millis, R. Mueller, and B.I. Shraiman, Phys. Rev. B\n54, 5405 (1996).\n47C. Aruta et al., Phys. Rev. B 80, 140405(R) (2009)." }, { "title": "1507.04490v1.Antiferromagentic_resonance_detected_by_DC_voltages_in_MnF__2__Pt_bilayers.pdf", "content": "Antiferromagentic resonance detected by DC voltages in MnF 2/Pt bilayers\nPhilipp Ross,1, 2,a)Michael Schreier,1, 3,b)Johannes Lotze,1, 3Hans Huebl,1, 4Rudolf Gross,1, 3, 4and\nSebastian T. B. Goennenwein1, 4\n1)Walther-Meißner-Institut, Bayerische Akademie der Wissenschaften, Garching, Germany\n2)Present address: London Centre for Nanotechnology, University College London, London, UK\n3)Physik-Department, Technische Universit¨ at M¨ unchen, Garching, Germany\n4)Nanosystems Initiative Munich, Munich, Germany\n(Dated: February 22, 2018)\nWe performed coplanar waveguide-based broadband ferromagnetic resonance experiments on the antiferro-\nmagnetic insulator MnF 2, while simultaneously recording the DC voltage arising in a thin platinum film\ndeposited onto the MnF 2. The antiferromagnetic resonance is clearly reflected in both the transmission\nthrough the waveguide as well as the DC voltage in the Pt strip. The DC voltage remains largely unaffected\nby field reversal and thus presumably stems from microwave rectification and/or heating effects. However, we\nidentify a small magnetic field orientation dependent contribution, compatible with antiferromagnetic spin\npumping theory.\nI. INTRODUCTION\nFerromagnetic resonance1describes the collective, res-\nonant excitation of magnetic moments in a ferromag-\nnet and is a powerful tool to investigate a specimen’s\nmagnetic properties. In ferromagnet/normal metal het-\nerostructures, ferromagnetic resonance furthermore rep-\nresents an effective source of pure spin currents2, i.e. an-\ngular momentum transport without an associated charge\nflow. In the so-called spin pumping process, the nor-\nmal metal acts as a spin sink, accepting non-equilibrium\nangular momentum from the resonantly driven magne-\ntization in the adjacent ferromagnet. If heavy metals\nwith large spin-orbit coupling are used as spin sinks, the\nspin current can be converted into an electrical current\nby means of the inverse spin Hall effect3,4and thus de-\ntected electrically. The electrically detected spin pump-\ning process has been investigated in a large variety of\nmagnetic materials5–7, including magnetic insulators8,9\nsuch as yttrium iron garnet. The magnetic materials\nhereby are either ferro- or ferrimagnetic which brings\nalong some limitations for experiments and potential ap-\nplications. At moderate (few Tesla) magnetic fields ferro-\nand ferrimagnets are limited to magnetic resonance fre-\nquencies of a few to a few ten GHz and their finite mag-\nnetization makes them susceptible to e.g. unintended\nstray fields. While some of these features can be useful\nfor certain applications, faster dynamics and resilience\nto external factors are often priority design goals. An-\ntiferromagnets offer magnetization dynamics at several\nhundred GHz even at zero external field, have negligi-\nble stray fields and are virtually immune to unintended\nchanges in their magnetization state. These qualities at\nthe same time, however, make it challenging to investi-\ngate antiferromagnetic dynamics. One option is to apply\nlarge magnetic fields to tune the antiferromagnetic reso-\na)Electronic mail: philipp.ross.13@ucl.ac.uk\nb)Electronic mail: michael.schreier@wmi.badw.denance frequency to a frequency range covered by conven-\ntional magnetic resonance spectrometers. Furthermore,\nthe properties of antiferromagnets with respect to spin\ncurrents are largely unexplored. However, recent exper-\niments demonstrate spin current transmission through\nantiferromagnets10and theory predicts11that antiferro-\nmagnets also could be useful as spin current source.\nHere we present antiferromagnetic resonance experi-\nments on the antiferromagnetic insulator MnF 212close\nto the spin-flop transition. We simultaneously record the\nmicrowave absorption signal and the DC voltages aris-\ning along a thin Pt strip deposited on top of the MnF 2\ncrystal. While the antiferromagnetic resonance is clearly\nreflected in the DC voltage, the voltage does not invert\nsign under reversal of the external field as one would\nexpect for conventional spin pumping. This indicates\nthe presence of additional DC voltage generation mecha-\nnisms, potentially rectification by means of the spin Hall\nmagnetoresistance, and/or thermal effects. A small but\nfinite change in the amplitude of the DC voltage under\nfield reversal may be attributed to spin pumping from\nMnF 2into Pt.\nII. THEORY\nWe consider a simple antiferromagnet with two mag-\nnetic sublattices and a parallel aligned external magnetic\nfieldH0. As detailed in Refs. 13 and 14, the fundamen-\ntal (exchange) magnetic resonance mode of such a system\nallows for the two solutions\nω±\nres=γµ0/bracketleftBig\n±H0+/radicalbig\nHA(2HE+HA)/bracketrightBig\n. (1)\nThe above, however, only holds for external fields H0be-\nlow a critical field HC, which marks the so-called spin-flop\ntransition. When H0> H cboth sublattice magnetiza-\ntions rotate to a new configuration nearly perpendicular\nto the magnetic field and bend towards H0with increas-\ning field strength. As the two sublattices are degenerate\nin this configuration, the magnetic resonance responsearXiv:1507.04490v1 [cond-mat.mtrl-sci] 16 Jul 20152\n0 3 6 9 120200400600\nμ0H0(T)f(GHz)ωres+\nωres-\nωresμ0HC\nFIG. 1. Antiferromagnetic resonance in MnF 2for external\nfields along the magnetic easy-axis. Below the spin-flop field\nµ0HC≈9.5 T magnetic resonance can occur for two different\nfrequencies, one increasing ( ω+\nres) and one decreasing ( ω−\nres) as\nH0increases.ω−\nreseventually approaches zero as H0→HC.\nAboveHCthe two collapse to a single resonance line.\nreduces to a single line at the frequency15\nωres=γµ0/radicalBig\nH2\n0−HA(2HE+HA). (2)\nIn Eqs. (1) and (2) above γis the gyromagnetic ratio,\nµ0is the vacuum permeability, HEis the exchange field,\nHAis an uniaxial anisotropy field and HHH0is assumed\nparallel toHHHA. MnF 2is well approximated in this pic-\nture usingµ0HE≈53 T,µ0HA≈0.85 T,µ0HC≈9.5 T\n(Refs 12, 16, and 17) and γ= 0.92γe. Whileω+\nresis typi-\ncally too large to be accessible using standard microwave\nequipment, ω−\nresapproaches zero frequency when H0is\nclose toHC(cf.Fig. 1). Even a small misalignment\nθH(cf.Fig. 2) between HHH0andHHHA, however, will shift\nω−\nres(H0≈HC) upwards by approximately 4 GHz/deg\nagain.\nIn (ferro-)magnetic resonance the magnetic material\n“pumps” a spin current into an adjacent normal metal,\nwhich thereby provides an additional damping channel\nfor magnetization excitation in the magnetic material2.\nIn electrically detected spin pumping this spin current\ngenerates an electric field EISHalong the normal metal\ndue to the inverse spin Hall effect3,4. AsEISHdepends\non the relative orientation between voltage-taps and spin\ncurrent spin polarization (which itself is linked to the\nmagnetization orientation of the ferromagnet), the volt-\nageVISHmeasured in experiment is antisymmetric with\nrespect to a reversal of the external magnetic field, i.e.\nVISH(+H0) =−VISH(−H0)6,18. Theory suggests11that\nthe same should also hold for antiferromagnetic reso-\nnance experiments, with the spin current spin polariza-\ntion determined by the antiferromagnet’s N´ eel-vector.\nhrf\ngold wir e\ncontact padsCPW with\nSU-8 isolationlayerMnF 2 with Pt strip \nMini-SMP\nconnectorsFIG. 2. The MnF 2/Pt sample is placed on a coplanar waveg-\nuide structure capped with an insulating resist layer. The Pt\nstrip is contacted via Au wires to enable detection of DC volt-\nages. The external magnetic field is applied along the mag-\nnetic easy-axis of MnF 2, however, the mechanical mounting\nlimits alignment precision to a few degrees.\nIII. SAMPLE AND EXPERIMENTAL SETUP\nFor the experiments we use a commercial MnF 2single\ncrystals in the shape of a 3 mm ×1 mm×0.5 mm cuboid\nwhose magnetic easy-axis points along the 0 .5 mm side.\nAfter polishing the sample a 3 mm ×250µm, 7 nm thick\nPt film is deposited on the 3 mm ×0.5 mm side of the sam-\nple by electron beam evaporation using a shadow mask.\nThe sample is then placed on a coplanar waveguide such\nthat the microwave field is perpendicular to the magnetic\neasy-axis (Fig. 2) and thereby able to effectively excite\nthe antiferromagnetic resonance12. The waveguide in-\ncluding the sample mounted on top is placed in a liquid\nHelium magnet cryostat which provides a magnetic field\nalong the magnetic easy-axis of the MnF 2crystal. All ex-\nperiments are performed in He exchange gas at pressures\nof few mbar and a temperature of 4 K /lessmuchTN´ eel\nMnF 2≈67 K19.\nIV. RESULTS\nFigure 3b shows a false color plot of the transmission\n|˜S21|2through the coplanar waveguide for microwave fre-\nquencies 15 .5 GHz≤f≤19.5 GHz and external field\nstrengths of 9 T≤µ0H0≤9.9 T. The data were recorded\nusing a vector network analyzer by performing magnetic\nfield sweeps at fixed frequencies ( Pmw= 11 dBm). For\nclarity, the transmission at an off-resonant offset value\n|S21(H0= 9 T)|2(Fig. 3b) [|S21(H0=−9 T)|2, Fig. 3a]\nhas been subtracted from each field sweep. The ap-\nproximately parabolic feature with reduced transmis-\nsion corresponds to the excitation of the antiferromag-\nnetic resonance by the microwave field. As discussed in\nSec. II the resonance frequency does not drop to zero at\nH0=HC≈9.5 T but remains finite due to a small mis-\nalignment of about θH≈4.21◦, inadvertently introduced3\n|S\n212\n0\n-0.04-0.03-0.02-0.01\n-9.9 -9.6 -9.3 -9.016171819\nμ0H0(T)f(GHz)a\n|S\n212\n0\n-0.04-0.03-0.02-0.01\n9.0 9.3 9.6 9.916171819\nμ0H0(T)f(GHz)b\nVDC(µV)\n-10+1\n-9.9 -9.6 -9.3 -9.016171819\nμ0H0(T)f(GHz)c\nVDC(µV)\n-10+1\n9.0 9.3 9.6 9.916171819\nμ0H0(T)f(GHz)d\nFIG. 3. Transmission change through the coplanar waveguide in vicinity of the critical field ( a,b). The antiferromagnetic\nresonance along with some spin wave modes are well reflected as a drop in the transmitted power. Only minor differences exist\nbetween the data for positive ( b) and negative ( a) applied external magnetic fields. The dashed line is a simulation of the\nantiferromagnetic resonance frequency for a misalignment between uniaxial anisotropy axis and the external magnetic field of\n4.21◦. The simultaneously recorded DC voltage along the Pt film ( c,d) closely resembles the transmission data. Differences\nbetween positive and negative applied external fields are, again, small but a slight change in amplitude is visible.\nwhen mounting the sample. The dashed line indicat-\ning the antiferromagnetic resonance frequency is calcu-\nlated based on an extension of the model by Skrotskii\nand Kurbatov20to two sublattices. When using the mis-\nalignment given above and the parameter set introduced\nin Sec. II the experimental data are well reproduced.\nUpon closer inspection at least two spin wave modes21\ncan be observed. Figure 3a shows data recoded in the\nsame fashion for fields applied antiparallel to HHHA, i.e.\n−9 T≥H0≥−9.9 T. As expected this spectrum closely\nmirrors the one for positive external magnetic fields, with\nonly minor differences which we will discuss in Sec. V.\nFigures 3c and 3d show the DC voltage across the\nPt strip recorded simultaneously with the transmission\ndata. As with the transmission data an offset voltage\nVDC(H0=−9 T) (Fig. 3c) [ VDC(H0= 9 T), Fig. 3d] was\nsubtracted from the raw data for clarity. Clearly, the an-\ntiferromagnetic resonance is also visible in VDC, as are theindividual spinwaves. In contrast to |˜S21|2, however, the\ndifferences between VDC(H0>0) andVDC(H0<0) are\nmore pronounced. Although the sign of the DC voltage\nremains unchanged under field reversal the magnitude of\ntheVDC(H0<0) data is slightly less than that of the\nVDC(H0>0) data. It is worth noting that, in the imme-\ndiate vicinity of HC, neither transmitted power nor DC\nvoltage seem to be significantly affected by the spin-flop\ntransition.\nV. DISCUSSION\nThe character of the data prevent a clear identification\nof the mechanism(s) responsible for the DC voltage\ngeneration. It is apparent from Figs. 3c and 3d that VDC\nis largely unaffected by the field reversal. Since these\nvoltages are not compatible with the characteristic sign4\nΔ|S\n212\n0\n-0.04-0.03-0.02-0.01\n9.0 9.3 9.6 9.916171819\n|μ0H0| (T)f(GHz)a\nΔVDC(µV)\n-10+1\n9.0 9.3 9.6 9.916171819\n|μ0H0| (T)f(GHz)b\nFIG. 4. ( a) Transmitted power difference between negative\nand positive applied external magnetic fields. Both positive\nand negative values are recorded, however, the average value\nis less than zero. This indicates more efficient microwave ab-\nsorption at negative fields. ( b) DC voltage difference between\nnegative and positive applied external magnetic fields. A clear\nbias towards positive values is observed, signifying the larger\nmagnitude of the DC voltage for positive external magnetic\nfields.\ninversion upon field reversal expected from electrically\ndetected spin pumping in ferromagnets, a different\nmechanism must be dominating.\nMicrowave absorption by the antiferromagnet is\ntied to the antiferromagnetic resonance ( cf. Figs. 3a\nand 3b) and can heat the sample substantially22,23. If\nthe heating is nonuniform across the sample a thermal\nvoltage can be generated in the Pt strip24. At the\nincident microwave power of 11 dBm and given the\nlarge thermal conductivity of the sample25,26, however,\nit appears unrealistic to expect nonuniform heating of\nmore than few ten millikelvins between the two contacts\nseparated by a distance of 300 µm. On the other hand,\nwith the Seebeck coefficient for the Pt-Au-wire-junction\nwell below 1 µV/K27,28atT= 4 K, it would take a\ntemperature difference between the two voltage contacts\nof more than 1 K to account for all of the magnetic fieldorientation independent offset.\nThe spin Seebeck effect29is another possible spurious\neffect and may occur if the temperature profile has a\nfinite slope along the MnF 2/Pt interface normal. The\ndetection of the effect, however, relies on the same\nmechanism as spin pumping and should therefore also\ndepend on the sign of the external magnetic field. Also,\nas supported by first experiments30, the spin Seebeck\neffect is predicted31to vanish in antiferromagnets.\nThus, the spin Seebeck effect is an unlikely candidate to\naccount for the dominant features detected in the DC\nvoltage.\nFilms in which the electrical resistivity is tied to the\nmagnetization orientation can show a so-called mi-\ncrowave rectification effect32. In these materials, under\nmagnetic resonance, the resistivity oscillates at the same\nfrequency as the external driving field. Simultaneously\nthe microwave driving field can induce an oscillating\ncharge current in the material. The product of the\nhigh frequency resistance and the charge current results\nin a DC voltage. While MnF 2itself is insulating, the\nresistivity of the Pt film may be linked to the sublattice\nmagnetization orientations in the MnF 2by means of the\nspin Hall magnetoresistance33. Unfortunately spin Hall\nmagnetoresistance theory34can not readily be applied to\nantiferromagnets. The field dependence of the resulting\nantiferromagnetic spin Hall magnetoresistance mediated\nrectification effect35(if any) is thus unclear as of writing\nthis manuscript.\nTo analyze the difference between positive and\nnegative applied external field in more detail we\ncompute the microwave transmission difference\n∆|˜S21|2=|˜S21(H0<0)|2−|˜S21(H0>0)|2(Fig. 4a).\nWe find that/angbracketleft∆|˜S21|2/angbracketright<0, where/angbracketleft···/angbracketright denotes the\naverage over all data points. In contrast, the analogously\ncomputed ∆ VDC=VDC(H0<0)−VDC(H0>0)\n(Fig. 4b) yields/angbracketleft∆VDC/angbracketright>0. Hence, although slightly\nmore power was absorbed at negative fields, the DC\nvoltage magnitude was notably larger for positive fields.\nFrom its symmetry properties under field inversion, this\nfield dependent voltage ∆ VDCis consistent with the\nspin pumping mechanism outlined in Sec. II. If spin\npumping indeed is at the origin of the observed voltage\ndifference, this would correspond to a contribution of\naboutVSP(±H0)≈∓100 nV which is about a fifth to\na tenth of the total DC voltage observed. Note also\nthat electrically detected spin pumping experiments\ntypically yield DC voltages in the range of a few 10 nV\nto a few 10 µV, such that the magnitude of the voltage\nalso appears reasonable.\nVI. CONCLUSION\nIn summary we investigated DC voltages arising when\nexciting antiferomagnetic resonance in MnF 2/Pt bilayer\nsamples. Placing the sample on a coplanar waveguide,5\nthe antiferromagnetic resonance below and above MnF 2’s\nspin-flop transition is observed in both microwave trans-\nmission through the waveguide as well as in the DC volt-\nage detected along the Pt strip. Since the DC voltage\ndoes not invert its sign under field reversal, and thus does\nnot exhibit the fingerprint of electrically detected spin\npumping, the DC signal must be dominated by other\nprocesses. We envisage rectification or, somewhat less\nlikely, thermal effects potentially result in the DC volt-\nage signature. In detail, we note the presence of a small\ndifference in VDCbetween positive and negative applied\nexternal magnetic fields, which may indicate antiferro-\nmagnetic spin pumping. However, further studies are\nrequired to substantiate this conjecture.\nACKNOWLEDGMENTS\nThe authors gratefully acknowledge financial support\nby the DFG via SPP 1538 “Spin Caloric Transport”\n(project GO 944/4-2).\nREFERENCES\n1C. Kittel, Phys. Rev. 73, 155 (1948).\n2Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys. Rev.\nLett. 88, 117601 (2002).\n3M. I. D’yakonov and V. I. Perel’, JETP Lett. 13, 467 (1971).\n4J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).\n5M. V. Costache, M. Sladkov, S. M. Watts, C. H. van der Wal,\nand B. J. van Wees, Phys. Rev. Lett. 97, 216603 (2006).\n6O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D.\nBader, and A. Hoffmann, Phys. Rev. Lett. 104, 046601 (2010).\n7F. D. Czeschka, L. Dreher, M. S. Brandt, M. Weiler, M. Altham-\nmer, I.-M. Imort, G. Reiss, A. Thomas, W. Schoch, W. Limmer,\nH. Huebl, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett.\n107, 046601 (2011).\n8C. W. Sandweg, Y. Kajiwara, K. Ando, E. Saitoh, and B. Hille-\nbrands, Appl. Phys. Lett. 97, 252504 (2010).\n9B. Heinrich, C. Burrowes, E. Montoya, B. Kardasz, E. Girt, Y.-\nY. Song, Y. Sun, and M. Wu, Phys. Rev. Lett. 107, 066604\n(2011).\n10T. Moriyama, S. Takei, M. Nagata, Y. Yoshimura, N. Matsuzaki,\nT. Terashima, Y. Tserkovnyak, and T. Ono, Appl. Phys. Lett.\n106, 162406 (2015).11R. Cheng, J. Xiao, Q. Niu, and A. Brataas, Phys. Rev. Lett.\n113, 057601 (2014).\n12M. Hagiwara, K. Katsumata, H. Yamaguchi, M. Tokunaga, I. Ya-\nmada, M. Gross, and P. Goy, Int J Infrared Millimeter Waves\n20, 617 (1999).\n13C. Kittel, Phys. Rev. 82, 565 (1951).\n14F. Keffer and C. Kittel, Phys. Rev. 85, 329 (1952).\n15T. Nagamiya, K. Yosida, and R. Kubo, Adv. Phys. 4, 1 (1955).\n16F. M. Johnson and A. H. Nethercot, Phys. Rev. 114, 705 (1959).\n17J. P. Kotthaus and V. Jaccarino, Phys. Rev. Lett. 28, 1649\n(1972).\n18M. Schreier, G. E. W. Bauer, V. I. Vasyuchka, J. Flipse, K. ichi\nUchida, J. Lotze, V. Lauer, A. V. Chumak, A. A. Serga, S. Dai-\nmon, T. Kikkawa, E. Saitoh, B. J. van Wees, B. Hillebrands,\nR. Gross, and S. T. B. Goennenwein, J. Phys. D: Appl. Phys.\n48, 025001 (2015).\n19C. Kittel, Introduction to Solid State Physics (Wiley, 2004).\n20G. Skrotskii and L. Kurbatov, in Ferromagnetic Resonance ,\nedited by S. Vonsovskii (Pergamon, 1966) pp. 12 – 77.\n21C. Herring and C. Kittel, Phys. Rev. 81, 869 (1951).\n22F. Sakran, A. Copty, M. Golosovsky, D. Davidov, and P. Monod,\nAppl. Phys. Lett. 84, 4499 (2004).\n23N. Yoshikawa and T. Kato, J. Phys. D: Appl. Phys. 43, 425403\n(2010).\n24T. J. Seebeck, Abhandlungen der K¨ oniglichen Akademie der Wis-\nsenschaften zu Berlin 1822-1823 , 265 (1825).\n25G. A. Slack, Phys. Rev. 122, 1451 (1961).\n26W. Haynes, CRC Handbook of Chemistry and Physics, 95th Edi-\ntion (CRC Press, 2014).\n27P. Blood and D. Grieg, J. Phys. F 2, 79 (1972).\n28J. Kopp, J. Phys. F 5, 1211 (1975).\n29K. Uchida, H. Adachi, T. Ota, H. Nakayama, S. Maekawa, and\nE. Saitoh, Appl. Phys. Lett. 97, 172505 (2010).\n30S. Gepr¨ ags, A. Kehlberger, T. Schulz, C. Mix, F. Della Co-\nletta, S. Meyer, A. Kamra, M. Althammer, G. Jakob, H. Huebl,\nR. Gross, S. T. B. Goennenwein, and M. Kl¨ aui, ArXiv e-prints\n(2014), arXiv:1405.4971 [cond-mat.mes-hall].\n31Y. Ohnuma, H. Adachi, E. Saitoh, and S. Maekawa, Phys. Rev.\nB87, 014423 (2013).\n32H. J. Juretschke, J. Appl. Phys. 31, 1401 (1960).\n33H. Nakayama, M. Althammer, Y.-T. Chen, K. Uchida, Y. Kaji-\nwara, D. Kikuchi, T. Ohtani, S. Gepr¨ ags, M. Opel, S. Takahashi,\nR. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh,\nPhys. Rev. Lett. 110, 206601 (2013).\n34Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B.\nGoennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87,\n144411 (2013).\n35R. Iguchi, K. Sato, D. Hirobe, S. Daimon, and E. Saitoh, Appl.\nPhys. Expr. 7, 013003 (2014)." }, { "title": "1509.02836v1.Inertial_terms_to_magnetization_dynamics_in_ferromagnetic_thin_films.pdf", "content": "arXiv:1509.02836v1 [cond-mat.mtrl-sci] 9 Sep 2015Inertial terms to magnetization dynamics in ferromagnetic thin\nfilms\nY. Li1,4, A.-L. Barra2, S. Auffret3,4,5, U. Ebels3,4,5, and W. E. Bailey1∗\n1Materials Science & Engineering, Dept. of Applied Physics &\nApplied Mathematics, Columbia University, New York NY 10027, USA\n2Laboratoire National des Champs Magn´ etiques Intenses,\nCNRS/UJF/UPS/INSA, 38042 Grenoble Cedex, France\n3Univ. Grenoble Alpes, INAC-SPINTEC, F-38000 Grenoble, France\n4CNRS, SPINTEC, F-38000 Grenoble, France\n5CEA, INAC-SPINTEC, F-38000 Grenoble, France\n∗Correspondence to: web54@columbia.edu\n(Dated: July 9, 2021)\nAbstract\nInertial magnetization dynamics have been predicted at ult rahigh speeds, or frequencies ap-\nproachingtheenergy relaxation scaleof electrons, inferr omagnetic metals. Hereweidentifyinertial\nterms to magnetization dynamics in thin Ni 79Fe21and Co films near room temperature. Effective\nmagnetic fields measured in high-frequency ferromagnetic r esonance (115-345 GHz) show an addi-\ntional stiffening term which is quadratic in frequency and ∼80 mT at the high frequency limit of\nour experiment. Our results extend understanding of magnet ization dynamics at sub-picosecond\ntime scales.\n1The magnetization M(t) in ferromagnetic materials is generally understood to evolve\nwithout memory of its prior motion. As described by the Landau-Lifs hitz (LL) equation[1],\nmagnetization dynamics dM(t)/dtcan be written in terms of the magnetization M(t) alone,\nexcluding any temporal derivatives dnM(t)/dtn. Magnetization is then ’inertialess’: it re-\nsponds to a step magnetic field H(t) with an instantaneous change in speed and with infinite\nacceleration.\nAs pointed out by Ciornai et al.and F¨ ahnle et al.[2–4], followed by other theoreticians[5,\n6], the absence of inertia is questionable for magnetization dynamics at very high frequen-\ncies. The high frequency behavior, >100 GHz, becomes relevant in ultrafast ( 100 GHz to reflect the scattering of six repeats. The dashed lines a re low-\nfrequency linear extrapolations from α0. Compared with extrapolated linewidths from the\nlow-frequency ∆ H1/2(ω), we observe reduced linewidths at high frequencies (115-345 GHz ).\nNo explicit prediction has been made for the effect of rotational iner tia on the linewidth in\nRef. [2, 12], but the observed behavior matches well with the predic tion of linear inertial\nterms in Ref. [16–18]. The solid curves are fits to the form:\nµ0∆H1/2=µ0∆H0+2α0ω\nγ1\n1+ω2τ2(3)\ntaking the α0from the low-frequency linewidths. The fitted τare listed in Table 1. The\nvalues of τextracted from µ0∆H1/2are close but slightly smaller than from µ0Hres. We\nnote here that the linewidth measurements are less precise than µ0Hresand more sensitive\nthan resonance fields to various sources of inhomogeneities in the s amples.\nAt room temperature, the relaxation rate of Bloch states (1 /τB) is determined by the\nelectronscatteringwithphononsandimpurities. Inthissense, τBissimilar totheremagneti-\nzation time τEin the ultrafast demagnetization experiments[19–21], where optica lly excited\nelectrons also relax through electron-phonon and electron-impur ity interactions. In the limit\nof zero laser fluence, nonlinear effects due to high occupation numb er of excited states are\nreduced[21, 22]. The zero-fluence τEhas been reported to be 0.2-0.25 ps for Py 10-30 nm[22]\nand∼0.4 ps for Co 15 nm[21], close to the value of τin this work.\nAt high frequencies the (nonmagnetic) skin depths δsin the ferromagnetic ���lms become\nsmaller and the enhanced eddy current effect may influence the res onance field[23]. The res-\nonance field will be enhanced by ∼µ0Ms/δ4\nsk4\n0, wherek0is the lowest-energy wavenumber\ndetermined by the surface anisotropy. Because δ2\ns∝1/ω, the resonance field enhancement\nis proportional to ω2and may influence the quadratic term in Eq. (2). Our calculations\nshow that this term is negligible, about 0.4 mT for Py 30 nm and 0.09 mT fo r Co 30 nm,\ncompared with the observed effects up to 80 mT (See the Supplemen tal Information for\ndetails).\nWe do not believe that interfacial effects, including spin pumping, play an important role\n8in the observed high-frequency behavior. Both Gilbert damping α0and the two inertial\ndynamics lifetimes τ(µ0Hres) andτ(µ0∆H1/2) in Table 1 show little thickness dependence\nfor either Py or Co, indicating that bulk relaxation is dominant. The we ak thickness de-\npendence of α0is consistent with the very low spin pumping effect of Py/Ta identified in\nRef. [24], in any case negligible for 30 nm films and without quadratic fre quency depen-\ndence. Theoretical predictions of resonance shifts from imaginar y spin mixing conductance\nare three orders of magnitude lower than observed here[25]. Only τ(µ0Hres) measured from\nresonance shifts for 6 nm films, not matched in τ(µ0∆H1/2) measured from linewidths, differ\nsignificantly. This enhancement may be structural in origin.\nA technological implication of our results is that effective field require ments for preces-\nsional switching will be reducedas switching times in magnetic storage decrease into the\nfew-picosecond range. In this sense, the inertial dynamics ease u ltrafast switching, if the\nbehaviors of Py and Co are representative of other metallic ferrom agnets. The effective field\nreduction, here up to 80 mT in Co, is not small in an absolute sense, an d might according\nto Eq. (2) be enhanced significantly in ferromagnets with higher Gilbe rt damping. On the\nother hand, prior switching experiments with high-field relativistic ele ctron bunches seem\nto indicate that nonlinear damping increases effective field requireme nts by a rather larger\namount for large-angle dynamics in CoCrPt[8], underscoring the utilit y of HF-FMR to iden-\ntify the inertial effect.\nIn summary, we identify a novel term to magnetization dynamics in th e ferromagnetic\nmetal films Py and Co which is quadratic in frequency and becomes sign ificant above 100\nGHz. Thetermstiffensthefrequency, aidingappliedfieldsindriving ult rafastmagnetization\nmotion. The behavior is best explained by dynamics retarded throug h a finite Bloch-state\nrelaxation time τBas proposed in Refs. [3, 4, 6]. Extracted relaxation times are 0.1-0.2 ps\nfor Py and 0.2-0.4 ps for Co, close to the remagnetization times meas ured in optical pump-\nprobe demagnetization experiments. Our findings extend underst anding of magnetization\ndynamics at picosecond time scales and may open up new possibilities fo r high-speed inertial\nswitching in ferromagnetic materials, previously demonstrated only in antiferromagnets[10].\nWe thank M. F¨ ahnle, J. E. Wegrowe and J. Fransson for discussion s and X. Yang for\nsuggestions on statistical analysis. We acknowledge fundings by th e EC through CRONOS\n(N◦280829), NSF-DMR-1411160, and the Chair of Excellence Program of the Nanosciences\nFoundation in Grenoble France for support.\n9[1] Eq. (1) is equivalent to the Landau-Lifshitz-Gilbert (L LG) equation with renormalized gyro-\nmagnetic ratio γ. See: T. L. Gilbert, IEEE Trans. Magn. 6, 3443 (2004).\n[2] M.-C. Ciornei, J. M. Rubi and J.-E. Wegrowe, Phys. Rev. B 83, 020410(R) (2011).\n[3] M. F¨ ahnle, D. Steiauf and C. Illg, Phys. Rev. B 84, 172403 (2011).\n[4] M. F¨ ahnle, D. Steiauf and C. Illg, Phys. Rev. B 88, 219905(E) (2013).\n[5] D. B¨ ottcher and J. Henk, Phys. Rev. B 86, 020404(R) (2012).\n[6] S. Bhattacharjee, L. Nordstr¨ om and J. Fransson, Phys. Rev. Lett. 108, 057204 (2012).\n[7] C. H. Back, R. Allenspach, W. Weber, S. S. P. Parkin, D. Wel ler, E. L. Garwin and H. C.\nSiegmann, Science285, 864 (1999).\n[8] I. Tudosa, C. Stamm, A. B. Kashuba, F. King, H. C. Siegmann , J. St¨ ohr, G. Ju, B. Lu and\nD. Weller, Nature428, 831 (2004).\n[9] C. D. Stanciu, F. Hansteen, A. V. Kimel, A. Kirilyuk, A. Ts ukamoto, A. Itoh and Th. Rasing,\nPhys. Rev. Lett. 99, 047601 (2007).\n[10] A. V. Kimel, B. A. Ivanov, R. V. Pisarev, P. A. Usachev, A. Kirilyuk and Th. Rasing, Nature\nPhys.5, 727 (2009).\n[11] S. Mangin, M. Gottwald, C.-H. Lambert, D. Steil, V. Uhl´ ıˇ r, L. Pang, M. Hehn, S. Alebrand,\nM. Cinchetti, G. Malinowski, Y. Fainman, M. Aeschlimann and E. E. Fullerton, Nature Mater.\n13, 286 (2014).\n[12] J.-E Wegrowe and M.-C. Ciornei, Am. J. Phys. 80, 607 (2012).\n[13] Y. Li and W. E. Bailey, arXiv1401.6467\n[14] A.-L. Barra, A. K. Hassan, A. Janoschka, C. L. Schmidt an d V. Sch¨ unemann, Appl. Magn.\nReson.30, 385 (2006).\n[15] Paramagnetic markers used: BDPA complex with benzene( 1:1) free radical for low-frequency\nFMR,geff=2.0025; MnO diluted in the diamagnetic host MgO for high-fr equency FMR,\ngeff=2.00101\n[16] V. Kambersk´ y, Can. J. Phys. 48, 2906 (1970).\n[17] V. Kambersk´ y, Phys. Rev. B 76, 134416 (2007).\n[18] A. M. Clogston, Bell Syst. Tech. J. 34, 739 (1955).\n[19] E. Beaurepaire, J.-C. Merle, A. Daunois and J.-Y. Bigot ,Phys. Rev. Lett. 76, 4250 (1996).\n10[20] B. Koopmans, J. J. M. Ruigrok, F. Dalla Longa and W. J. M. d e Jonge, Phys. Rev. Lett. 95,\n267207 (2005).\n[21] B. Koopmans, G. Malinowski, F. Dalla Longa, D. Steiauf, M. F¨ ahnle, T. Roth, M. Cinchetti\nand M. Aeschlimann, Nature Mater. 9, 259 (2009).\n[22] C. A. C Bosco, A. Azevedo and L. H. Acioli, Appl. Phys. Lett. 83, 1767 (2003).\n[23] P. Pincus, Phys. Rev. 118, 658 (1960).\n[24] S. Mizukami, Y. Ando and T. Miyazaki, J. Magn. Magn. Mater. 226, 1640 (2001).\n[25] M. Zwierzycki, Y. Tserkovnyak, P. J. Kelly, A. Brataas a nd G. E. W. Bauer, Phys. Rev. B\n71, 064420 (2005).\n11" }, { "title": "1502.04128v1.Antiferromagnet_Mediated_Spin_Transfer_Between_Metal_and_Ferromagnet.pdf", "content": "Antiferromagnet-Mediated Spin Transfer Between Metal and Ferromagnet\nSo Takei,1Takahiro Moriyama,2Teruo Ono,2and Yaroslav Tserkovnyak1\n1Department of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\n2Institute for Chemical Research, Kyoto University, 611-0011 Uji, Kyoto, Japan\n(Dated: June 8, 2022)\nWe develop a theory for spin transported by coherent N ´eel dynamics through an antiferromagnetic insulator\ncoupled to a ferromagnetic insulator on one side and a current-carrying normal metal with strong spin-orbit cou-\npling on the other. The ferromagnet is considered within the mono-domain limit and we assume its coupling to\nthe local antiferromagnet N ´eel order at the ferromagnet jantiferromagnet interface through exchange coupling.\nCoupling between the charge current and the local N ´eel order at the other interface is described using spin\nHall phenomenology. Spin transport through the antiferromagnet, assumed to possess an easy-axis magnetic\nanisotropy, is solved within the adiabatic approximation and the e \u000bect of spin current flowing into the ferromag-\nnet on its resonance linewidth is evaluated. Onsager reciprocity is used to evaluate the inverse spin Hall voltage\ngenerated across the metal by a dynamic ferromagnet as a function the antiferromagnet thickness.\nPACS numbers: 72.25.Mk, 75.47.-m, 75.50.Ee, 76.50. +g\nSpintronics of antiferromagnets (AFs), where AFs take on\nthe role of the central active component, is identified as one of\nthe most important emerging topics in the field of magnetism\ntoday [1]. Robustness to magnetic perturbations due to their\ntotal magnetic compensation, as well as characteristic dynam-\nical scale in the THz range may render AFs advantageous over\nferromagnets (Fs) for spintronics device applications. In ad-\ndition, recent works on AFs have shown that the important\nphenomena responsible for the success of F-based spintron-\nics also have AF counterparts, giving added impetus for AF-\nbased spintronics research. Indeed, giant magnetoresistance\nand current-induced torques [2], anisotropic magnetoresis-\ntance [3] and spin superfluidity [4], as well as current-induced\ndomain wall motion [5] and coupled dynamics between con-\nduction electrons and background magnetic texture [6], are all\nshown to be possible in AFs as well.\nAn important aspect of AF-based spintronics is the use of\nAFs as a medium to transport spin angular momentum. Spin\ntransfer through AFs has been the focus of several recent ex-\nperimental endeavors. Both Hahn et al . [7] and Wang et\nal. [8] demonstrated spin transport through an AF insulator,\nNiO, using an YIG jNiOjPt heterostructure (YIG standing for\nthe insulating ferrimagnet yttrium iron garnet). Inverse spin\nHall signal showed robust spin pumping from YIG into Pt\neven in the presence of the intervening NiO, suggesting e \u000e-\ncient spin transport through the AF. More recently, Moriyama\net al . used spin-torque ferromagnetic resonance (ST-FMR)\nto demonstrate the propagation of spin excitations through a\nmetallic AF, IrMn, using a Pt jIrMnjCoFeB trilayer [9] as well\nas NiO using a PtjNiOjFeNi trilayer [10]. Spin current injected\nfrom the Pt was shown to change the FMR linewidth, also sug-\ngesting the transfer of spin angular momentum through the\ncentral AF. Given the rising interest in AF spintronics and\nthe recent experimental focus, a theoretical account of spin\ntransport through an experimentally relevant normal metal\n(N)jAFjF trilayer is highly desirable.\nIn this Letter, we develop a general phenomenology for spin\ntransport through an AF by collective N ´eel order parameter\nx=0x=LxzNAFFS(t)✓⌧\"n(x)j(j,\")(S,⌧)JhJssttihJsGiNAFF(#,↵\"#)#⇤effective junctionVISHE`\n↵0hJsspihJsexiFIG. 1. Normal-metal (N) jantiferromagnet (AF) jferromagnet (F) tri-\nlayer considered in this work. N sustains a dc charge current jand\nF is described by a time-dependent macrospin S(t). Spin transfer\nhJs\nexioccurs via the exchange coupling Jat the AFjF interface, while\nspin transfer across the AF jN interface has a spin transfer torque con-\ntributionhJs\nstti(proportional to the e \u000bective interfacial spin Hall an-\ngle#) and a spin pumping contribution hJs\nspi(proportional to the\ninterfacial spin-mixing conductance \u000b\"#). The AF Gilbert damping,\nparametrized by \u000b0, leads to the loss of spin current hJs\nGiin the AF\nbulk. The central AF can be thought of as an e \u000bective interface that\ncouples jandSwith an e \u000bective spin Hall angle #\u0003.\ndynamics, focusing on an N jAFjF trilayer relevant for both the\nspin-pumping /inverse spin Hall as well as the ST-FMR exper-\niments mentioned above (see Fig. 1). Spin Hall phenomenol-\nogy, applicable to a wide range of di \u000berent AFjN interfaces\nobeying certain structural /crystalline symmetries, is utilized\nto model the spin transfer at the AF jN interface, while the ex-\nchange coupling is assumed at the AF jF interface. As one of\nthe main achievements of this work we develop a simple “cir-\ncuit” model, a pictorial visualization of spin flow, that allows\none to keep track of spin transfer through various parts of the\nheterostructure (see bottom half of Fig. 1). From the circuit\nmodel, we see that spin is both injected into (i.e., hJs\nstti) andarXiv:1502.04128v1 [cond-mat.mes-hall] 13 Feb 20152\nejected (i.e.,hJs\nspi) out of the AF at the AF jN interface due to\nspin Hall /spin-torque e \u000bects and spin pumping, respectively.\nThe collective N ´eel dynamics leads to Gilbert damping and to\nthe loss of spin current (i.e., hJs\nGi) within the AF bulk, and the\nexchange coupling at the AF jF interface leads to spin transfer\n(i.e.,hJs\nexi) across the interface. We first study how spin trans-\nport through the AF modifies the linewidth at FMR, akin to the\nST-FMR [9, 10]. The FMR linewidth is quantified in terms of\nthe e\u000bective spin Hall angle and spin-mixing conductance at\nthe AFjN interface, the exchange coupling at the AF jF inter-\nface, as well as AF Gilbert damping. We show that linewidths,\nmeasured for di \u000berent electrical currents in N and AF thick-\nnesses, can be used to extract the e \u000bective spin Hall angle\nand spin-mixing conductance at the N jAF interface, as well as\nthe bulk Gilbert damping. By invoking Onsager reciprocity,\nwe also make connections with the inverse spin Hall experi-\nments [7, 8] and compute the inverse spin Hall voltage gener-\nated across N arising as a result of a dynamic F macrospin.\nAs shown in Fig. 1, an insulating AF is attached on one side\nto a mono-domain F and on the other to a paramagnetic N with\nstrong spin-orbit coupling. The N and F sustain dc charge\ncurrent density jand a time-dependent macrospin S(t), re-\nspectively. We consider a bipartite AF, which can be charac-\nterized by two hydrodynamic variables, n(x;t) andm(x;t),\nparametrizing the staggered (N ´eel) and smooth (magnetic)\ncomponents of the spins, respectively [11]. We assume easy-\naxis magnetic anisotropy along the zaxis in the AF, as well\nas full translational and rotational symmetries in the yzplanes\nso that our treatment essentially reduces to a one-dimensional\nproblem that depends only on the coordinate x. The free en-\nergyFfor the AF and its coupling to the F reads\nF=ZL\n0dx(A\n2[@xn(x)]2+m(x)2\n2\u001f\u0000\u0014\n2nz(x)2)\n\u0000JS\u0001n(L);(1)\nwhere Aand\u001fare the N ´eel order sti \u000bness and spin suscepti-\nbility, respectively, \u0014the uniaxial anisotropy parameter, and J\nthe exchange coupling between AF and F.\nThe Landau-Lifshitz-Gilbert dynamics in the bulk AF cor-\nresponding to Eq. (1) can be written as\ns( ˙n+\u000bn\u0002˙m)=\u001f\u00001m\u0002n; (2)\ns( ˙m+\u000bm\u0002˙m+\u000b0n\u0002˙n)=n\u0002(A@2\nxn+\u0014nzez);(3)\nwhere\u000band\u000b0are (independent) Gilbert damping param-\neters and sis the roughly saturated spin density (per unit\nlength) [12]. In the limit of slow dynamics (i.e., \u001c\u001d~=Eex,\nwhere\u001cis the time scale for AF dynamics and Eexis the\nAF exchange energy scale) and strong local N ´eel order (i.e.,\njm(x;t)j\u001c1), one may first use Eq. (2) to solve for m, insert\nthe solution into Eq. (3), and arrive at the following dynamics\nfor the AF N ´eel vector,\n\u001fs2n\u0002¨n+s\u000b0n\u0002˙n\u0019An\u0002@2\nxn+\u0014n\u0002nzez:(4)The AF spins are excited by the current and the dynamic\nmacrospin. The e \u000bects of these external perturbations are lo-\ncalized at the interfaces and thus enter the AF dynamics as\nboundary conditions. At the AF jN interface, spin currents\narise via spin transfer torque and spin-pumping, both of which\ncan be accounted for using spin Hall phenomenology [13].\nBased on structural symmetries at the interface, spin Hall phe-\nnomenology allows us to write down a general expression for\nthe spin transfer torque that applies to a variety of F- and AF-\nbased heterostructures with di \u000berent microscopic details. In\nthe presence of full translational and rotational symmetries\nin the yzplane and with the breaking of reflection symmetry\nalong the xaxis, there are two contributions to the spin current\n(integrated over the interface area) flowing into AF [13]\nJs\nl=[#n\u0002(ex\u0002j)\u0002n\u0000~\u000b\"#n\u0002˙n]jx=0\u0011Js\nstt\u0000Js\nsp;(5)\nwhere the first term is the so-called spin Hall-like (dissipative)\ncontribution and the second term describes spin-pumping.\nThe coe \u000ecient#is proportional to the (tangent of the) ef-\nfective spin Hall angle at the AF jN interface [13]; although #\ncan, in general, depend on the orientation of n, we will treat\nit as a constant here. We will also disregard any anisotropies\nof\u000b\"#with respect to the orientations of nand ˙n, assuming\nthat the exchange energy scale at the interface dominates over\nthe energy scale of spin-orbit interactions.\nWhile the coe \u000ecient\u000b\"#is proportional to the real part of\nthe (generally complex) spin-mixing conductance g\"#for the\nAFjN interface [13], its imaginary part gives rise to a term\n/˙nin Eq. (5). The assumed structural symmetry, in prin-\nciple, also allows for the so-called field-like (reactive) con-\ntribution given by \u0011(ex\u0002j)\u0002n. Assuming the magnitudes\nof all four terms are small (as compared top\nA\u0014to be pre-\ncise) we will compute the FMR linewidth to linear-order in\nthe coe \u000ecients#,\u0011andg\"#. Since the terms that do not ap-\npear in Eq. (5) are non-dissipative terms, they do not enter the\nlinewidth at linear-order in these coe \u000ecients. We will, there-\nfore, drop these contributions from the following discussion\nand only consider the two terms in Eq. (5).\nThe exchange coupling J[given by the last term in Eq.(1)]\nleads to the following expression for spin current flowing\nacross the AFjF interface,\nJs\nr\u0011Js\nex=@F\n@S\u0002S=JS\u0002njx=L: (6)\nThe spin current in the AF bulk can be read o \u000bfrom Eq. (3)\n(dropping the Gilbert damping and anisotropy terms) and the\nresulting continuity equation [i.e., s˙m=\u0000@x(\u0000An\u0002@xn)], so\nthatJs\nAF(x)=\u0000An(x)\u0002@xn(x). The continuity of spin current\nacross each interface then leads to the boundary conditions\nJs\nl=Js\nAF(x=0);Js\nr=Js\nAF(x=L): (7)\nThe dynamic N ´eel texturen(x;t) can be obtained using the\nlow-frequency (adiabatic) approximation, valid in the regime\n\n\u001c\n0, where \nand\n0are the FMR and the AF resonance3\n(a)(b)L/\u0000\u0000↵F\u0000↵F\nFIG. 2. The interfacial contribution \u000e\u000b(i)\nF(dashed lines), the bulk\ncontribution \u000e\u000b(b)\nF(dotted lines) and the total contribution \u000e\u000bF(solid\nlines) to the extrinsic FMR linewidth ( Sset to unity) are plotted as a\nfunction of the (normalized) system size L=\u0015. We fix the following\nparameters: \u0011=1,#jy=b0=0:01 and\u000b\"#=0:01. Two regimes\nare considered for the AF Gilbert damping ˜ \u000b: (a) the strong damping\nregime ˜\u000b=0:2; and (b) the weak damping regime ˜ \u000b=0:01 (see text\nfor more details).\nfrequencies, respectively. Within this approximation, the AF\nN´eel texture is first solved for an arbitrary static S, the result\ndenoted byn(0)(x;S). SinceS(t) varies su \u000eciently slowly in\ntime compared to the characteristic AF time scale, the N ´eel\ntexture in the adiabatic limit will arrange itself into the static\nconfiguration corresponding to S(t) at every moment in time\nand is given by n(x;t)\u0019n(0)[x;S(t)]\u0011n(0)(x;t). The above\ncalculation does not account for spin current losses due to the\nAF dynamics (i.e., spin-pumping at the AF jN interface and\nGilbert damping in the AF bulk). Taking these losses into ac-\ncount up to linear-order corrections to the adiabatic result, the\nspin currenthJs\nexientering F, time-averaged over a cycle of\nFMR precession (the angle brackets h\u0001\u0001\u0001ihereafter represent-\ning time-average over a cycle of FMR precession), is given\nbyhJs\nexi=hJs\nstti\u0000hJs\nspi\u0000hJs\nGi(c.f. Fig. 1), where the spin-\ntransfer torque contribution is given by inserting the adiabatic\nresult for the N ´eel texture into Eq. (6)\nhJs\nstti=JhS(t)\u0002n(0)(L;t)i; (8)\nand the loss terms read\nhJs\nspi=~\u000b\"#hn(0)(0;t)\u0002˙n(0)(0;t)i;\nhJs\nGi=s\u000b0ZL\n0dxhn(0)(x;t)\u0002˙n(0)(x;t)i:(9)\nThe first term in Eq. (9) describes (time-averaged) spin current\nlost due to spin pumping at the AF jN interface and the second\nterm corresponds to Gilbert damping in the AF bulk.\nAn analytical result for the FMR linewidth can be ob-tained if we consider small deviations of S(t) away from\nthezaxis (parallel to the static FMR field and the AF easy-\naxis); we take j=jyey+jzezand assumejjjto be weak\nsuch that a linear-response treatment is su \u000ecient. In this\ncase, the N ´eel unit vector nshould not deviate far from the\nzaxis and we may evaluate the above results with respect\nto small transverse fluctuations, i.e., S(t)\u0019S[sx(t);sy(t);1]\nandn(x;t)\u0019[nx(x;t);ny(x;t);1] withjsx(t)j;jsy(t)j\u001c 1 and\njnx(x;t)j;jny(x;t)j \u001c 1. Within this treatment, the trans-\nverse components n?=(nx;ny)Tobey\u001fs2¨n?+s\u000b˙n?=\nA@2\nxn?\u0000\u0014n?[c.f. Eq. (4)], and n(0)(x;t) has the form\nn(0)\u0019ez+f(x)[ez\u0002S(t)]\u0002ez+g(x)ez\u0002S(t)+h(x)ex;(10)\nwhere the functions f(x) and g(x) (to linear-order in the cur-\nrent) are given by\nf(x)=1\nScoshx\n\u0015\ncoshL\n\u0015+1\n\u0011sinhL\n\u0015; (11)\ng(x)=1\nSsinhL\u0000x\n\u0015+1\n\u0011coshL\u0000x\n\u0015\u0010\ncoshL\n\u0015+1\n\u0011sinhL\n\u0015\u00112#jyp\nA\u0014; (12)\nandh(x)/#jzis not explicitly shown here since this term will\nnot contribute to the linewidth within the current theoretical\ntreatment. Here, \u0011\u0011JS=p\nA\u0014, and\u0015\u0011pA=\u0014is the AF\nhealing length.\nThe spin current Js\nexentering F modifies the F dynamics as\n~˙S=b\u0002S\u0000~\u000bF\nSS\u0002˙S+Js\nex; (13)\nwhere\u000bFis the intrinsic Gilbert damping parameter in F and\nb=\u0000b0ezis the static FMR field (in units of energy). Insert-\ning Eq. (10) into Eqs. (8) and (9) and performing the time-\naverage over the last two terms in Eq. (13), the full FMR\nlinewidth can be read o \u000bdirectly by summing the coe \u000ecients\nappearing in front of hS\u0002˙Si. The total Gilbert damping pa-\nrameter is then given by \u000b0\nF=\u000bF+\u000e\u000b(i)\nF+\u000e\u000b(b)\nF\u0011\u000bF+\u000e\u000bF,\nwhere the extrinsic contribution \u000e\u000bFhas the interfacial contri-\nbution\u000e\u000b(i)\nFand the AF bulk contribution \u000e\u000b(b)\nF:\n\u000e\u000b(i)\nF=1\nS\u0010#jy\nb0+\u000b\"#\u0011\n\u0010\ncoshL\n\u0015+1\n\u0011sinhL\n\u0015\u00112; (14)\n\u000e\u000b(b)\nF=˜\u000b\nSL\n\u0015+1\n2sinh2L\n\u0015\u0010\ncoshL\n\u0015+1\n\u0011sinhL\n\u0015\u00112; (15)\nwhere ˜\u000b=s\u000b0\u0015=2~. The former originates from spin injection\nand spin-pumping at the AF jN interface while the latter from\nGilbert damping in the AF bulk. Eqs. (14) and (15) constitute\nthe main result of this work.\nAs seen from Eqs. (14) and (15), the healing length \u0015sets\nthe distance over which spin propagation decays inside the\nAF. The healing length is determined from the slope of the\nlinewidth vs. jycurves for various thicknesses Land by ex-4\ntracting the decay length. It is important to note that the cur-\nrent theory only considers spin transport mediated by coherent\nN´eel dynamics, and does not take account of spin transported\nby incoherent thermal magnons. The latter contribution is\nsuppressed at su \u000eciently low temperatures by some power of\nthe ratio T=TN, where TNis the N ´eel ordering temperature\nof the AF, and since magnon-mediated transport is expected\nto decay over the spin di \u000busion length \u0015sd, it is strongly sup-\npressed for \u0015sd\u001cL.\nOnce the AF healing length is known, the parameters #,\u000b\"#\nand ˜\u000bcan be extracted by measuring the FMR linewidth for\nvarious jyandL. While the e \u000bective spin Hall angle #can be\nobtained from the slope of a linewidth vs. jycurve, the Gilbert\ndamping parameter ˜ \u000bcan be extracted in the regime L\u001d\u0015,\nin which the linewidth depends only on ˜ \u000b(see Fig. 2),\n\u000e\u000bFL\n\u0015!1\n!˜\u000b\nS \u0011\n1+\u0011!2\n\u0011\u000e\u000b1\nF: (16)\nForL\u001c\u0015, we expand \u000e\u000bFto linear order in L=\u0015,\n\u000e\u000bF\u00191\nS #jy\nb0+\u000b\"#!\n+1\nS\"\n˜\u000b\u00002\n\u0011 #jy\nb0+\u000b\"#!#L\n\u0015\n\u0011c0+c1L\n\u0015;(17)\nfrom which we see that \u000b\"#can be extracted at zero current\n(i.e., jy=0) and measuring the linewidth for L\u001c\u0015.\nThe extrinsic linewidth exhibits qualitatively di \u000berent be-\nhavior depending on the relative magnitudes of the bulk and\nthe interfacial contributions (see Fig. 2). For strong Gilbert\ndamping ( c1>0) the bulk damping in the AF dominates over\nthe interface e \u000bects and the linewidth grows initially as Lin-\ncreases, saturating eventually as L=\u0015!1 [see Fig. 2(a)]. In\nthe limit of weak Gilbert damping [see Fig. 2(b)], i.e., c1<0\n(and\u000e\u000b1\nFωrf\nγ. We will present details of\nthe analysis elsewhere [23].\nWe assign the lower field feature in the FMRFM spectra shown in Fig. 1 t o the reso-\nnance contributions originating from the localized FMR excitations spatially confined ap-\nproximately to region I of the sample. In this region, the resonance occurs at lower values\nofHextthan that of the main peak (Fig. 1, dotted line). The frequency shif t oflocalized\nFMR is determined by two factors: (1) the strength of the tip field Htipat the sample\nsurface, and (2) the effect of mode confinement to the spatial re gion I with characteristic\ndimensions defined by the tip-sample distance, which further increa ses thelocalmode fre-\nquency relative to the bulk resonance by a value ∆ ωconf(r′). Both effects cause the local\nresonance to occur at the external field value that is lowerthan that of the bulk resonance\n4by the amount ∆ Hext≈ −Htip(r′)−∆ωconf(r′)/γ. From numerical simulations [23], for a\nconfinement within a disc of radius r′∼10µm, ∆ωconf(r′)/γ≈30 Oe, which combined with\nthe estimated value of Htip(r′)≈20 Oe results in a total shift consistent with experimental\nfindings. Both the tip field Htipand ∆ωconf(r′) decrease as the probe magnet is retracted\naway from the film surface, and local modes merge into the main reso nance (Fig. 1, dotted\nline).\nThe non-lorentzian and broad shape of the signal possibly indicates the presence of mul-\ntiple modes contributing to the resonance. While numerical simulation s are required to\ndetermine the possibility of such modes in particular geometry/mate rials, we believe that\ntheir appearance is generic in thin-film soft magnets and is induced by the local field inho-\nmogeneity. We will present the detailed theoretical analysis elsewhe re[23].\nThe normalized FMRFM spatial force map obtained from the uniform F MR mode mod-\nified by the tip field is shown in Fig. 2. The semicircular region of the plot w here the\ntip-sample interaction force is set to zero corresponds to region I . In the lower panel of\nFig. 2 we show the intensity of the main resonance signal as a functio n of the tip-sample\ndistance and compare simulations with the experiment. The dashed lin e shows the expected\nforce signal for the uniform FMR mode, not modified by the tip field. I n this case there is\nno force exerted by a uniformly magnetized infinite film on a spherical probe tip.\nIn conclusion, we conducted FMRFM experiments in a thin permalloy film . We per-\nformed quantitative analysis of the force exerted by the fundame ntal mode and observed\nlocally excited FMR. We conducted simulations and determined two dist inctive regions of\nthe sample contributing to the FMRFM spectra. We find clear evidenc e for local modifica-\ntion of the FMR mode structure by the probe tip providing insight into the interaction of\nthe probe tip with ferromagnetic samples in FMRFM.\nThe work performed at Los Alamos National Laboratory was suppo rted by the US De-\npartment of Energy, Center for Integrated Nanotechnologies, contract W-7405-ENG-36 at\nLos Alamos National Laboratory and contract DE-AC04-94AL850 00 at Sandia National\nLaboratories. The work at Ohio State University was supported by the US Department\nof Energy through grant DE-FG02-03ER46054. The work at Nava l Research Laboratory\n(NRL) was supported by the Office of Naval Research through the Institute for Nanoscience\nat NRL.\n5[1] J. A. Sidles, Appl. Phys. Lett. 58, 2854 (1991)\n[2] D. Rugar, C. S. Yannoni, and J. A. Sidles, Nature 360, 563 (1993)\n[3] D. Rugar and O. Z¨ uger and S. Hoen and C. S. Yannoni and H. M. Vieth and R. D. Kendrick,\nScience264, 1560 (1994)\n[4] Z. Zhang, P. C. Hammel, and P. E. Wigen, Appl. Phys. Lett. 68, 2005 (1996)\n[5] V. V. Naletov, V. Charbois, O. Klein, and C. Fermon Appl. P hys. Lett. 83, 3132 (2003)\n[6] V. Charbois, V. V. Naletov, J. Ben Youssef, and O. Klein, A ppl. Phys. Lett. 80, 4795 (2002)\n[7] R. Urban, A. Putilin, P. E. Wigen, S.-H. Liou, M. C. Cross, P. C. Hammel, and M. L. Roukes,\nPhys. Rev. B 73, 212410 (2006)\n[8] D. Rugar, R. Budakian, H. J. Mamin, and W. Chui, Nature 430, 329 (2004)\n[9] O. Z¨ uger, and D. Rugar, Appl. Phys. Lett. 63, 2496 (1993)\n[10] A. Suter, D. V. Pelekhov, M. L. Roukes, and P. C. Hammel, J . Magn. Res. 154, 210 (2002)\n[11] www.attocube.com, models ANPx(z) 100/LIN\n[12] E. Nazaretski, T. Mewes, D. Pelekhov, P. C. Hammel, and R . Movshovich, AIP Conf. Proc.\n850, 1641 (2006)\n[13] www.magnequench.com, partciles type: MQP-S-11-9\n[14] E. Nazaretski, J. D. Thompson, D. V. Pelekhov, T. Mewes, P. E. Wigen, J. Kim, M. Zalalut-\ndinov, J. W. Baldwin, B. Houston, P. C. Hammel, and R. Movshov ich, J. Magn. Magn. Mat.\n(2006), doi:10.1016/j.jmmm.2006.10.994\n[15] I. Dorofeyev, H. Fuchs, G. Wenning, and B. Gotsmann, Phy s. Rev. Lett. 83, 2402 (1999)\n[16] E. Nazaretski, J. D. Thompson, M. Zalalutdinov, J. W. Ba ldwin, B. Houston, T. Mewes, D.\nV. Pelekhov, P. Wigen, P. C. Hammel, and R. Movshovich, J. App l. Phys. MS JR06-3902R\n(accepted)\n[17] T. Mewes, J. Kim, D. V. Pelekhov, G. N. Kakazei, P. E. Wige n, S. Batra, and P. C. Hammel,\nPhys. Rev. B 74, 144424 (2006)\n[18] Z. Frait, Physica 86-88B, 1241 (1977)\n[19] J. D. Jackson Classical Electrodynamics 3rdedition, Wiley, New York, 1999\n[20] C. Herring, and C. Kittel, Phys. Rev. 81, 869 (1951)\n[21] M. Sparks B. R. Tittmann, J. E. Mee, and C. Newkirk, J. App l. Phys.40, 1518 (1969)\n6[22] T. L. Gilbert, Phys. Rev. 100, 1243 (1955); L. D. Landau, E. M. Lifshitz, and L. P. Pitaevsk i,\nStatistical Physics, Part 2 (Pergamon, Oxford, 1980), 3rd ed.\n[23] D.V. Pelekhov, unpublished\n7Figure Caption\nFigure 1: Evolution of the FMRFM signals as a function of the probe-fi lm spacing.\nDotted line indicates the position of the main resonance peak, indepe ndent of the probe-\nfilm distance. Arrows mark the onset of the lower field resonance fe ature. Experimental\nparameters: ωrf/2π=9.55 GHz, T = 11 K\nFigure2: Upper panel: forcemapoftheFMRFMprobe-filminteractio n. Theforceacting\non a cantilever due to an elementary ring-shaped area is calculated a s a function of radius of\nthe ring andthe cantilever-film spacing. The forcemap is normalized t o a maximum positive\nforce value at each probe-film distance. Force contribution from r egion I of the sample is set\nto zero because the fundamental resonant mode which occurs at Hresdoes not significantly\npenetrate into region I where Hz\ntot>ωrf\nγ. The inset shows schematically the probe-sample\narrangement andtwo regionsofthesamplecontributing totheFMR FMsignal. Lower panel:\nintegrated probe-film interaction force as a function of the cantile ver-film spacing for the\nmain resonance peak at Hext∼14.54 kOe. Solid symbols represent experimental points and\nsolid line is the result of calculations based on the exclusion of region I. The dashed line\nshows the force if region I were included in integration. Two insets sh ow SEM micrographs\nof the cantilever tip.\n8102\n101\n102\n101\nExternal□field□[kOe]Signal□force□[10 N]-16\n100\n14 15 14.8 14.6 14.4 14.214.0 m /c1091009.0 m/c109103\n102\n101\n1005.5 m/c109103\n102\n101\n1002.5 m/c1092.5 m/c109\nFIG. 1:\nFIG. 2:\n9" }, { "title": "1605.07850v2.Exploring_the_ferromagnetic_behaviour_of_a_repulsive_Fermi_gas_via_spin_dynamics.pdf", "content": "Exploring the ferromagnetic behaviour of a repulsive Fermi gas via spin dynamics\nG. Valtolina,1;2;3F. Scazza,1;2A. Amico,1;2A. Burchianti,1;2\nA. Recati,4;5T. Enss,6M. Inguscio,1;2M. Zaccanti1;2and G. Roati1;2\n1INO-CNR, Via Nello Carrara 1, 50019 Sesto Fiorentino, Italy\n2LENS and Dipartimento di Fisica e Astronomia, Universit `a di Firenze, Via Nello Carrara 1, 50019 Sesto Fiorentino, Italy\n3Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy\n4INO-CNR BEC Center and Dipartimento di Fisica, Universit `a di Trento, 38123 Povo, Italy\n5Technische Universit ¨at M ¨unchen, James-Franck-Straße 1, 85748 Garching, Germany and\n6Universit ¨at Heidelberg, Philosophenweg 19, 69120 Heidelberg, Germany\nFerromagnetism is a manifestation of strong repulsive interactions between itinerant fermions in condensed\nmatter. Whether short-ranged repulsion alone is sufficient to stabilize ferromagnetic correlations in the absence\nof other effects, like peculiar band dispersions or orbital couplings, is however unclear. Here, we investigate\nferromagnetism in the minimal framework of an ultracold Fermi gas with short-range repulsive interactions\ntuned via a Feshbach resonance. While fermion pairing characterises the ground state, our experiments pro-\nvide signatures suggestive of a metastable Stoner-like ferromagnetic phase supported by strong repulsion in\nexcited scattering states. We probe the collective spin response of a two-spin mixture engineered in a magnetic\ndomain-wall-like configuration, and reveal a substantial increase of spin susceptibility while approaching a criti-\ncal repulsion strength. Beyond this value, we observe the emergence of a time-window of domain immiscibility,\nindicating the metastability of the initial ferromagnetic state. Our findings establish an important connection\nbetween dynamical and equilibrium properties of strongly-correlated Fermi gases, pointing to the existence of\na ferromagnetic instability.\nThe magnetic properties of a variety of quantum systems,\nranging from electrons in transition metals [1, 2] and normal\n3He liquids [3] to neutron and quark matter within the crust\nof neutron stars [4, 5], emerge from strong interactions be-\ntween itinerant fermions, i.e. not localized into a crystal lat-\ntice. Itinerant ferromagnetism can be qualitatively captured\nby an intuitively simple mean-field framework, as first en-\nvisioned by Stoner [6]: a free electron gas can become fer-\nromagnetic once a short-ranged screened Coulomb repulsion\nbetween oppositely oriented electron spins overcomes the ef-\nfect of Fermi pressure, which would favour a paramagnetic\nstate with no spin ordering. A sufficiently strong repulsion\npromotes the parallel alignment of magnetic moments, at the\nprice of an increased kinetic energy, making the paramagnetic\nstate unstable towards a ferromagnetic one. Stoner’s picture\nhas proven successful to qualitatively describe the phases of a\nwealth of electron systems. However, electrons in solids are\nsubject to various effects beyond short-range repulsion [1, 2],\nwhich are thought to play a role in promoting or suppress-\ning ferromagnetism. In particular, there exist materials where\na Stoner-type instability of the repulsive Fermi liquid state\nis altered by competing mechanisms. Notable examples are\nthe emergence of unconventional superconductivity adjoining\nferromagnetism in intermetallic compounds [7] and the non-\nFermi liquid behaviour of the paramagnetic state in itinerant-\nelectron ferromagnets [8].\nQuite paradoxically, even the paradigmatic scenario of an\nultracold atomic Fermi gas with short-ranged repulsive inter-\nactions entails a superfluid ground state of paired fermions,\nrather than a ferromagnetic one. This stems from the fact\nthat genuine zero-range repulsion, encoded in the s-wave scat-\ntering length athat can be adjusted via magnetic Feshbach\nresonances, necessarily requires an underlying attractive po-\ntential with a weakly bound molecular state [9]. Hence, the\nrepulsive Fermi gas corresponds to a metastable excited (up-per) energy branch of the many-body problem [10–12] (see\nFig. 1a). It is intrinsically unstable against pair formation via\ninelastic decay processes [11, 13, 14] that were found to be\nrapid with respect to the development of magnetic domains\n[13–15]. Nonetheless, theoretical approaches intentionally\nneglecting pairing, based on Landau’s Fermi liquid theory\nand quantum Monte Carlo (QMC) calculations [12, 16–21],\nconfirm Stoner’s ferromagnetic instability of a homogenous\nFermi gas driven only by short-range repulsion. In light of\nthe metastable nature of the repulsive branch, it is debated\nwhether such a system may exhibit ferromagnetic correlations\nat all, at least temporarily, and how the magnetic and transport\nproperties of the repulsive gas are affected by the large cou-\npling to lower-lying states.\nPrevious experiments with ultracold spin mixtures prepared\nin the paramagnetic phase and quenched to strong repul-\nsion [13, 22] found the Stoner instability to be precluded\nby the pairing one [15], concluding that a Fermi gas with\nstrong short-range repulsion does not undergo a ferromag-\nnetic phase transition [13]. On the other hand, a recent spec-\ntroscopic study [23] of highly-imbalanced repulsive mixtures\nin the polaronic regime [11, 12] revealed the Fermi liquid\nphase to become energetically and thermodynamically unsta-\nble beyond a critical repulsion strength, in quantitative agree-\nment with theoretical predictions for polarized repulsive gases\n[12, 19, 20, 24], while still featuring an unexpectedly long\nlifetime before decaying onto lower-lying states. This sup-\nports a scenario where the repulsive Fermi gas may exhibit\na transient time during which repulsion and ferromagnetic\ncorrelations dominate the evolution of the many-body system\nalso in the balanced case, for which a ferromagnetic instability\nis maximally favoured [20].\nIn this work we investigate the spin response and the sta-\nbility of a balanced spin \"–#mixture prepared in an ar-\ntificial magnetic domain-wall structure (see Fig. 1b). SucharXiv:1605.07850v2 [cond-mat.quant-gas] 22 Feb 20172\nz\nyx\nba\n-2 -1 0 1-2-1012ε/atom( a.u. )\n-1/kF a\n3\n03\nOptical density\nLR\nFIG. 1. Engineering a ferromagnetic state with an ultracold\natomic Fermi gas. a , A Fermi mixture with resonant short-range\ninteractions, parametrised by 1=kFa, features two distinct energy\nbranches, the lower (upper) being associated to a net interspecies\nattraction (repulsion). Ferromagnetism may develop at strong repul-\nsion along the upper branch, whose stability is limited due to en-\nhanced decay processes onto the lower branch. Depending on the\nspin imbalance, the lower branch corresponds either to a paired phase\nor to an attractive Fermi liquid. b, In our experiment we circumvent\nthe pairing instability by preparing an atomic Fermi gas in a fer-\nromagnetic domain-wall structure, probed via spin-selective in situ\nimaging (see lower image). The initial state is obtained by segregat-\ning the two spin components into two initially disconnected reser-\nvoirs at equilibrium by means of a thin optical barrier with a waist of\nabout 2\u0016m (sketched in green).\na fully-ferromagnetic initial configuration features a vanish-\ningly small\"–#density overlap in contrast to a paramagnetic\nstate [13, 22]. This greatly suppresses the effect of pairing\nprocesses, ensuring that the overall relaxation rate remains\nmuch slower than that given by the Fermi energy, thereby\nmaintaining the system on the upper branch for a compara-\ntively long time. We probe the ferromagnetic properties of the\nrepulsive Fermi gas through two distinct but interconnected\nmeasurements of spin dynamics at varying interaction. On the\none hand, by probing the spin-dipole mode [25], i.e. the out-\nof-phase relative oscillation of the approaching spin \"–#do-\nmains and observing its softening, we demonstrate an increase\nof the spin susceptibility as a function of repulsion strength,\nup to a sharp discontinuity in the spin response. On the other\nhand, by studying spin diffusion [28–30] at short and long\nevolution times we reveal the metastability of the initial ferro-\nmagnetic state and gain new insights on its relaxation mech-\nanisms [12, 13, 15]. In particular, our work shows that the\nshort-time collective dynamics of an artificially-created ferro-\nmagnetic state are governed by strong repulsion, before in-elastic decay to the attractive branch eventually leads to de-\nmagnetization.\nWe initially prepare a weakly interacting mixture of ultra-\ncold6Li atoms [31], equally populating the two lowest Zee-\nman states, hereafter denoted as j\"iandj#i. The atoms are\nheld in a cylindrical optical dipole trap with axial and radial\nfrequencies \u0017z'21Hz and\u0017?'265Hz, respectively. By\nadjusting the evaporation procedure we can tune the degree of\ndegeneracy from T=TF<0:1up to\u00181. HereTis the gas\ntemperature, while TFis the Fermi temperature of a single-\ncomponent Fermi gas of Natoms in a harmonic trap, given\nbykBTF=EF=h(6N\u0017z\u00172\n?)1=3, withhandkBdenot-\ning the Planck’s and Boltzmann’s constants. At a magnetic\nfield of about 1 G, where the magnetic moments of j\"iandj#i\nstates are opposite, the application of a magnetic field gradient\nallows us to spatially separate the two spin components along\nthe weak axis of the trap. Once the overlap between the two\nclouds is perfectly zero, we superimpose a 2 \u0016m thin optical\nrepulsive barrier as high as V0\u001810EFonto the center of\nthe harmonic potential, in order to split the trap into two inde-\npendent reservoirs [32], as sketched in Fig. 1b. We then adia-\nbatically turn off the magnetic field gradient and end up with\nall\"(#) fermions in the left (right) reservoir (see Methods\nand Supplementary Information for details). This creates two\nmacroscopic spin domains at rest, separated by a distance only\na few times wider than the mean interparticle spacing of the\ngas. Such configuration resembles the density distribution ex-\npected for a spin-mixture at full magnetization in an elongated\nharmonic trap, with a central domain wall of thickness around\nthe interparticle spacing, additionally surrounded by an unpo-\nlarized low-density shell on the cloud surface. From here, we\nlet the two spin components start interacting by removing the\noptical barrier. Before switching off the barrier, the interac-\ntion strength is adjusted by setting the magnetic field close to\nthe center of a broad \"–#Feshbach resonance located around\n832 Gauss [33].\nSpin response of a repulsive Fermi gas\nIn a first experiment the spin dynamics is triggered by\nabruptly switching off the barrier from its initial value on a\n\u0016s timescale. Owing to the small initial separation of about\n5\u0016m, the two spin clouds approach each other with small\nrelative momentum ~k\u001c~kF=p2mLiEF. We follow\nthe clouds dynamics by resonant in-situ absorption imaging,\nmonitoring the evolution of the two spin domains (see Meth-\nods and Supplementary Information). On top of an over-\nall slow drift, the relative distance between their centres of\nmassd(t) =z\"(t)\u0000z#(t)presents a small-amplitude out-of-\nphase oscillation, signalling the excitation of the spin-dipole\nmode (see Fig. 2a-c). The measurement of the spin-dipole\nfrequency[25, 26], next to the one of spin fluctuations [13, 27],\nis a powerful probe for disclosing the magnetic properties of\nharmonically trapped gases, connecting the local behaviour of\nthe spin susceptibility \u001fto a directly accessible quantity. Fol-\nlowing a sum-rule approach [25], the spin-dipole frequency\n\u0017SDis found to be:\n\u00172\nSD=N\"+N#\n4\u00192mLiR\ndrz2\u001f(n(~ r)); (1)3\nTime (ms)-1000100∆d(μm)\n0 20 40 60 80 100\nκF a = 0.04(μm)∆d05\n-5\n0 20 40 60 80 100κF a = 3.65\n0 20 40 60 80 10005\n-510(μm)∆dκF a = 1.02a\nb\nc\n d\nκF aνSD /νz\n0\n0 0.5 1 2 2.5 1.51.5\n0.52\n1\nνSD /νz\n123-1 0 1-1/(κF a)\ne\nFIG. 2. Spin response of a repulsive Fermi gas. a-c , After subtracting a slow exponential drift from the centre-of-mass distance d(t)between\nthe two spin clouds, the residual out-of-phase dynamics \u0001d(t)after sudden barrier removal is fitted to a damped sinusoidal function (dashed\nlines), from which \u0017SDis extracted for several interaction strengths. The bare trap oscillation is shown for comparison (dotted red line in\na). Shaded areas denote the standard confidence bands of the fits. Data points result from at least 5 independent measurements with error\nbars given by the standard error of the mean (s.e.m.) combined with the uncertainty on the subtracted exponential drift. d, The normalized\nspin-dipole frequency \u0017SD=\u0017zis plotted versus \u0014FaforT=TF= 0:12(2) (blue circles) and T=TF= 0:25(4) (purple squares), with error bars\nbeing 95 %confidence intervals of sinusoidal fits. A decrease of \u0017SDfollowed by a clear discontinuity is visible, suggesting the occurrence of a\nferromagnetic instability. Here, the shaded blue interval denotes the range of interaction strengths where the spin-dipole mode does not exhibit\na distinct single-frequency oscillatory behaviour. The dashed blue (violet) lines are the average \u0017SD=\u0017zmeasured beyond the discontinuity\npoint atT=TF'0:12(T=TF'0:25) up to unitarity. The solid (dashed) green lines are the T= 0 predictions from a sum-rule approach\nassuming 25 %(100%)\"–#spatial overlap (see Methods). e,\u0017SD=\u0017zmeasured atT=TF'0:12across the Feshbach resonance is shown for\nthe gas initially prepared in the upper branch (blue circles) or in the lower branch (red diamonds). The solid (dashed) red line represents the\nspin-dipole frequency prediction using the attractive gas spin susceptibility from Ref. 37 and setting a 20 %(100%)\"–#spatial overlap (see\nSupplementary Information). The dashed blue line coincides with the one shown in panel d.\nwhere\u001fdepends on the local density n(~ r)in the trap (see\nSupplementary Information for details). In particular, for a\nnon interacting gas \u0017SD=\u0017z. An increase of \u001fin the re-\npulsive Fermi liquid phase can therefore be distinctly iden-\ntified by\u0017SD< \u0017z, i.e. by a softening of the spin-dipole\nmode. Although \u001fis locally divergent if a ferromagnetic in-\nstability is reached at the trap center, \u0017SDcan only reach a\nnon-zero minimum [25] owing to the inhomogeneous density\nprofile of the trapped gas, whose outer low-density paramag-\nnetic region contributes with a small spin susceptibility. The\nmeasurement of this collective oscillation is extremely chal-\nlenging when starting from a paramagnetic configuration due\nto strong damping [34] and inelastic processes [11, 13, 23].\nHere, instead, where the two spin domains just partially over-\nlap over the timescale of the measurement, we are able to trace\na few oscillation periods of the spin-dipole mode from which\nwe extract\u0017SDthrough a fit to a damped sinusoidal function\n(see Supplementary Information). Performing such a mea-\nsurement at several magnetic field values, we obtain the trend\nof the spin-dipole frequency as a function of the repulsive in-\nteraction strength, displayed in Fig. 2d for two distinct tem-\nperatures. The interaction strength is described by the dimen-\nsionless parameter \u0014Fa, where\u0014F(and correspondingly \u000fF)\nis the average Fermi wave number (energy) weighted over the\ninitial density distribution close to the interface between the\ntwo domains (see Methods).Let us discuss here the results for the colder samples at\nT=TF= 0:12(2) (blue circles in Fig. 2d). By starting from\nthe weakly interacting regime, where \u0017SD'\u0017z(see Fig. 2a),\nan increase of the interspecies repulsion leads to a progres-\nsive reduction of the spin-dipole frequency, down to values as\nlow as\u0017SD'0:6\u0017zat about\u0014Fa'1(Fig. 2b). We find\nthe decrease of \u0017SDto be accompanied by a strong increase\nof the damping of the oscillations [34]. By further increasing\nthe interspecies repulsion, an abrupt change occurs in the spin\ndynamics: for \u0014Fa&1:1, the spin-dipole frequency sharply\njumps above the bare trap frequency, \u0017SD'1:70(4)\u0017z(see\nFig. 2c), while the damping of the oscillations is strongly re-\nduced. Once this narrow interaction region is crossed, a fur-\nther increase of \u0014Fadoes not produce any significant change,\nneither in the damping rate nor in \u0017SD. Higher tempera-\nture data exhibit a qualitatively similar trend, with the abrupt\nchange in\u0017SDoccurring at a higher \u0014Favalue.\nThe observed trend of \u0017SDup to\u0014Fa'1agrees with the\nmode-softening predicted by Eq. (1), reflecting a substantial\nincrease of spin susceptibility. This is further supported by\nthe good agreement of our low-temperature data with the spin-\ndipole frequency trend calculated in linear response [25] (see\ngreen lines in Fig. 2d), inserting into Eq. (1) the susceptibility\n\u001f(\u0014Fa)from QMC calculations for a T= 0homogeneous re-\npulsive Fermi gas [20] (see Methods). Furthermore, the value\nof\u0017SDfor\u0014Fa > 1.1, compatible with previous measure-4\nments at unitarity [28], compares well with the one expected\nfor two immiscible clouds of unpaired fermions bouncing off\neach other [35], as if one spin component acted like an im-\npenetrable potential barrier for the other. The increase of spin\nsusceptibility detected up to a critical repulsion strength to-\ngether with the discontinuity of the spin response frequency\nhint at a ferromagnetic instability located at \u0014Fa\u00191for\nT=TF'0:12.\nThe measured behaviour of \u0017SDcould not be explained if\npairing processes dominated the system dynamics. On the\nlower branch, owing to effective attraction, the spin suscep-\ntibility decreases monotonically while crossing the Feshbach\nresonance from a < 0toa > 0, vanishing completely for a\nBose-Einstein condensate (BEC) of pairs. This trend equally\noccurs in the attractive Fermi liquid state [36] as well as in\nthe superfluid or in the elusive pseudo-gap state [27, 37].\nTherefore, Eq. (1) implies a corresponding monotonic in-\ncrease of\u0017SDmoving towards the BEC limit. We experi-\nmentally demonstrate this behaviour by intentionally initial-\nizing the system in the lower branch (see Supplementary In-\nformation), observing an increase of \u0017SDeven above 2 \u0017zfor\n\u0014Fa > 0(see Fig. 2e). The spin response measured for the\nattractive gas is consistent with predictions from Eq. (1), ob-\ntained by inserting the spin susceptibility \u001frecently calcu-\nlated for a BEC-BCS crossover Fermi gas above the critical\ntemperature for superfluidity [37], in agreement with previous\nmeasurements [27]. In particular, it qualitatively differs from\nthe spin response at \u0014Fa>0shown in Fig. 2d, leading us to\nexclude that the latter arises from paired fermions in the inter-\nface region. On the other hand, to confirm that during the spin\ndynamics shown in Fig. 2a-c the upper branch is indeed pre-\ndominantly occupied, we estimate the molecular fraction un-\nder the same experimental conditions by an independent mea-\nsurement (see Supplementary Fig. S6). We detect a molecu-\nlar fraction not exceeding 15% after the first 50 ms of evolu-\ntion for any \u0014Favalue herein investigated, and below 5% for\n\u0014Fa\u001d1. Additionally, our findings seem incompatible with\na purely collisional picture [38] (see Supplementary Informa-\ntion), also considering the markedly distinct behaviour of \u0017SD\non the attractive branch shown in Fig. 2e.\nObservation of spin domain immiscibility\nIf the initially prepared fully ferromagnetic state was indefi-\nnitely stable above a critical \u0014Fa, the two spin domains would\nremain immiscible, i.e. spin diffusion would be impeded [34].\nTo investigate this aspect, we study spin diffusion in a sec-\nond set of measurements, exploring various interaction and\ntemperature regimes. We initialize the dynamics by adiabati-\ncally lowering the barrier height through a 30 ms linear ramp\nfromV0\u001810EFdown to 2EF, letting the two clouds slowly\napproach each other. Their relative distance is reduced from\nabout 5\u0016m down to 1 \u0016m, yet each spin domain remains con-\nfined within its own reservoir. At this point, we remove the\nbarrier in 5 ms and we monitor the subsequent evolution of\nthe relative population of the i=\";#component in the left and\nright reservoirs, Mi= (Ni;L\u0000Ni;R)=(Ni;L+Ni;R), from\nwhich we obtain the magnetization \u0001M= (M\"\u0000M#)=2.\nSince the distance between the two nearby cloud edges is ap-\nproximately equal to the local interparticle spacing at the in-terface between the two spin domains, this procedure does not\nexcite any detectable spin-dipole oscillation.\nFigure 3a shows the short-time evolution \u0001M(t)for differ-\nent interaction strengths: above a critical value of repulsion,\nafter an initial slight decrease of \u0001Mfrom 1 to about 0.92, we\nindeed observe a time window during which spin diffusion is\ncompletely arrested. We attribute the initial drop in magne-\ntization to spin transport within the low-density outer shells\nof the two spin clouds. The duration of the magnetization\n“plateau” is however finite, since the stability of our ferro-\nmagnetic state is limited by the intrinsic tendency of the sys-\ntem to relax from the excited upper branch onto lower-lying\nenergy states [12, 15] (see Fig. 1a). A thorough character-\nization of this interesting feature is summarized in Fig. 3b,\nwhere we plot the measured plateau duration \u001cpof constant\n\u0001Mas a function of 1=(\u0014Fa), for various temperatures. The\nvalue of\u001cpis determined through a piecewise linear fit to the\ndata, yielding zero when no noticeable halt of spin dynam-\nics is detected. A non-zero \u001cpis detected only above a crit-\nical\u0014Favalue and below a certain T=TF. Notably, finite\nplateaus appear above an interaction strength nearly coinci-\ndent with the one at which the spin-dipole mode frequency in\nFig. 2d reaches its minimum and exhibits an abrupt change\n(see Fig. 3c). Furthermore, by increasing \u0014Fa(increasing\nT=TF)\u001cpincreases (decreases), reaching its maximum at the\nunitary point 1=(\u0014Fa) = 0 . On the other hand, no dynamical\narrest is observed for T=TF\u00150:7. In addition, at low tem-\nperatures the trends for \u001cpand\u0017SD(see Fig. 3c and Fig. 2e)\nsuggest that we access the upper branch even within a narrow\n1=(\u0014Fa)<0region beyond unitarity, in accord with recent\npredictions [15].\nWe find the behaviour of the plateau duration for\n1=(\u0014Fa)>0to be captured (see curves in Fig. 3b) by a\nphenomenological model based only on the knowledge of the\nlifetime and energy spectrum of the upper and lower branches\nof the many-body system [24], calculated in the extremely po-\nlarized limit of one single \"(#) impurity embedded in a #(\")\nFermi gas [12, 24]. Based on such a description, the ferro-\nmagnetic state is destroyed by inelastic processes occurring\nat the interface between the two macroscopic spin domains:\nfermions of one kind, overcoming the surface tension associ-\nated with a domain wall, can deposit an overall excess energy\nthrough decay from the upper to the lower branch. Only af-\nter some time, once a sufficient energy has been released into\nthe system, the domain wall is melted and spin diffusion is\nestablished.\nUpon identifying for each temperature the lowest \u0014Fa\nvalue at which a non-zero \u001cpof steady magnetization is ob-\nserved, we delimit a region in the interaction-temperature\nplane where the ferromagnetic domains remain temporarily\nimmiscible, as displayed Fig. 3d. The critical interaction\nstrength for the onset of a non-zero \u001cpdisplays a non-linear\ndependence upon temperature, and by fitting the T=TF<0:3\ndata points with T=TF/((\u0014Fa)(T)\u0000(\u0014Fa)(0))\u000b, we ob-\ntain\u000b= 0:52(5) and(\u0014Fa)(0) = 0:80(9) . The fitted expo-\nnent matches within its uncertainty the value \u000b= 1=2ex-\npected from the low-temperature behaviour of a Fermi liq-\nuid exhibiting a magnetic instability (see Supplementary In-5\nτ P\n0 10 20 30 400.750.80.850.90.951\n \n \n \n \n \n∆M\nHold time (ms)1/κFa\n0.05\n0.45\n0.75\n1.3a\n2.5 2 1.5 1 0.5 0- 0.5 -105101520\nτP(ms)\n1/(κFa)0.06(2)\n0.12(2)\n0.31(4)\n0.53(6)T/TFb\n0 1 2 3 4 5 600.20.40.60.8T/T F\nκFaτ > 0Pτ = 0P\nd c\n3 2 1 0 -10.40.81.21.62 \n051015\n \nτP(ms)νSD /νz\n1/(κFa)\nFIG. 3. Metastability of a fully magnetized ultracold Fermi gas. a , Evolution of \u0001M(t)for different interactions at T=TF= 0:12(2) . For\n\u0014Fa\u00151:1a time window \u001cpof vanishing spin diffusion is detected. Error bars are the s.e.m. of 4-5 independent measurements. Data sets\nare artificially shifted by 0.05 along the y-axis from one another for clarity. b,\u001cpfrom piecewise fits to the data (dashed lines in a) is plotted\nat varying 1=\u0014Fafor various temperatures, with error bars being the fit standard uncertainty. Shaded curves are derived from the proposed\nmodel for domain-wall melting (see Methods), adjusting E+cwithin a 20% variation. c, Connection between \u001cp(red diamonds) and \u0017SD\n(blue circles) at T=TF= 0:12(2) . The grey region depicts the regime of \u001cp>0, which coincides within the experimental accuracy to the\nregion where \u0017SD'1:7. The dotted blue line denotes the prediction on \u0017SDfor the repulsive Fermi liquid (see Fig 2d). d, Metastability\n(\u001cp>0) region of the ferromagnetic state in the temperature-interaction plane. y-error bars denote the experimental uncertainty on T=TF,\nwhilex-error bars account for the uncertainty on estimating the critical \u0014Faat which a\u001cp>0is observed. The solid line is a power-law fit\n(see text) toT=TF<0:3points, extended over all values of \u0014Faas a dashed line. The black arrow marks the temperature for which \u001cp= 0\nat any\u0014Fa.\nformation). The extracted zero-temperature value (\u0014Fa)(0)\nis interestingly found in good agreement with the critical one\nobtained from repulsive QMC calculations [20, 21], and is sig-\nnificantly lower than Stoner’s mean-field criterion for an un-\npolarized gas [6, 12, 16–18, 20] \u0014Fa=\u0019=2. Notwithstand-\ning the metastable nature of the ferromagnetic state and the\ndynamical character of our study, our findings agree with the-\noretical expectations for a repulsive Fermi gas at equilibrium\nundergoing a ferromagnetic instability in the absence of pair-\ning. Moreover, the close correspondence between the trends\nof\u0017SDand\u001cp(see Fig. 3c) further suggests that the critical in-\nteraction strength for \u001cp>0corresponds to the one required\nfor the fully ferromagnetic state to be favoured.\nLong-time diffusive dynamics and spin drag coefficient\nOnce spin diffusion is established [28–30], the analysis of the\nlong-time evolution \u0001M(t)(or equivalently d(t)) within a\nsimple kinetic model (see Methods) allows us to determine\nalso the spin drag coefficient \u0000Sas a function of temperature\nand interaction. The results are displayed in Fig. 4. These are\ncompared with theoretical predictions for \u0000S, calculated fora single impurity diffusing in a homogeneous ideal Fermi gas\nwithin kinetic theory, accounting for scattering in all available\nstates within T-matrix approximation for the scattering cross\nsection (see Supplementary Information). The model is able\nto quantitatively reproduce the measured maximum of \u0000Sfor\nT=TF\u00150:3, that is the expected range of validity of the T-\nmatrix approximation, as well as the position of the maximum\nof\u0000Sat all temperatures (see Fig. 4). At low temperatures the\ndata sets exhibit a small but appreciable asymmetry around\nthe unitary point towards \u0014Fa>0. Such a feature, which dis-\nappears progressively as temperature is increased, highlights\nthe significant effect of collisions within the medium of sur-\nrounding particles on the dynamical properties of the diffusing\nquasi-particles (see Supplementary Information). A similar\nasymmetry in transport coefficients has already been reported\nfor the shear viscosity [39] and transverse spin diffusion [40],\nbut not in previous measurements of longitudinal spin diffu-\nsion [28].\nIn conclusion, we have probed the ferromagnetic behaviour\nof the repulsive Fermi gas by investigating spin dynamics in a\nresonantly interacting ultracold6Li spin mixture. Our findings6\n-1.0 -0.5 0 0.5 1.0 1.500.10.20.3 \n ћ ΓS / εF\n \n1/κFa-1.0 -0.5 0 0.5 1.0 1.50.00.10.20.30.4\n \n ћ ΓS / εFa\nbT/TF\n \n0.06(2) 0.12(2) 0.31(4) \n 0.53(6)\n 0.70(5)\nFIG. 4. Spin drag coefficient of a strongly interacting Fermi gas.\n~\u0000S=\u000fFis plotted as a function of 1=\u0014Fafora,0:31\u0014T=TF\u0014\n0:7, and b,T=TF\u00140:31. Experimental points are obtained by fit-\nting the dynamics at t >50ms to the solution of a diffusion model\n(see Methods). Error bars combine uncertainties of the fit and of\nour determination of \u000fF. Lines are predictions from a T-matrix ki-\nnetic theory (see Supplementary Information), assuming the nominal\ninitialT=TFand allowing a\u000620%temperature variation (shaded ar-\neas).\nindicate that the short-time evolution of the system is gov-\nerned by the effect of strong interparticle repulsion, and point\nto the existence of a Stoner-like ferromagnetic instability of\nthe zero-range repulsive Fermi gas, before relaxation onto at-\ntractive states alters irrevocably the nature of the many-body\nstate and its correlations. Furthermore, our study provides an\nimportant link between the dynamic response and the static\nproperties of a strongly correlated fermionic state. In the fu-\nture, our techniques may be exploited to probe spin transport\nand collective modes in different systems, such as repulsive\nand attractive Fermi gases with reduced dimensionality [41],\nand in the presence of weak optical lattices [42] and controlled\ndisorder [43].\nNote added – After submission of the manuscript, we became\naware of related theoretical work[44] by He et al. , report-\ning finite-temperature predictions qualitatively consistent with\nour findings.\nMETHODS\nExperimental protocols\nOur procedure to create weakly-interacting two-component Fermi\nmixtures of6Li atoms has been already described elsewhere[31]\n(see also Supplementary Information). The gas degeneracy parame-\nterT=TFis adjusted by exploiting the tunability of the collisional\nproperties of6Li mixtures[33] during evaporation. We typically\nend up with N\";#'5\u0002104atoms, confined in a cigar-shaped\nharmonic potential characterised by axial (radial) trap frequency of\n\u0017z= 21:0(1) Hz (\u0017?= 265(5) Hz). To spatially separate the twospin components, at a magnetic field of about 1 G, where the \"and\n#states possess equal but opposite magnetic moments, we turn on a\nmagnetic quadrupole gradient of about 1 G/cm along the weak trap\naxis, that pushes the two spin clouds towards opposite directions.\nOnce the overlap between the two components is zeroed, a repulsive\noptical potential centred at z= 0, and characterised by a short (long)\n1=e2waist ofwz= 2:0(2)\u0016m (wy= 840(30)\u0016m) is employed to\nconfine the two clouds into two disconnected reservoirs[32]. These\nare characterised by an axial oscillation frequency \u0017R= 1:78(5)\u0017z,\nand no appreciable particle tunnelling is detected over more than 2 s\nwith a barrier height of about 10 EF.\nTo excite the spin-dipole mode at a fixed Feshbach field, we\nabruptly switch off the barrier potential within less than 1 \u0016s. The\ntwo spin clouds, initially separated by a distance of about 5 \u0016m, start\nmoving one towards the other at a small relative velocity. From here\non, we monitor the dynamics of the relative distance between the\ncentres of mass of the two components. For each value of evolution\ntime, in two independent and successive experimental runs, we ac-\nquire two in-situ absorption images with two high-intensity optical\npulses, each of which is resonant only with one spin state. For each\npoint of interaction, temperature, and evolution time herein investi-\ngated, the relative distance between the two clouds is determined by\naveraging over at least five independent measurements. To investi-\ngate the spin diffusion of the two domains into each other, and in\nparticular to reveal the dynamical arrest of spin diffusion, we em-\nploy a different protocol turning off the repulsive barrier slowly in a\ntwo-stage sequence (see Supplementary Information for details).\nThe centre of mass of each cloud is equivalently determined either\nvia a Gaussian fit to the imaged density distribution (used for data\ndisplayed in Fig. 2a-c) or via the evaluation of the centre of mass zi\nof each sample i=\";#obtained via direct integration of the bare im-\nages. Once the spin dynamics is recorded, we fit d(t) = (z\"\u0000z#)(t)\nto an exponential decay. By subtracting the fitted exponential drift\nfromd(t), we extract the spin-dipole collective dynamics \u0001d(t),\nwhose frequency is then derived by means of a damped sinusoidal\nfunction (see Supplementary Information). Similarly, direct integra-\ntion of the images at each evolution time allows to determine the\nrelative population of the i=\";#component in the left and right\nreservoirs,Mi= (Ni;L\u0000Ni;R)=(Ni;L+Ni;R), from which we\nobtain the magnetization \u0001M= (M\"\u0000M#)=2. The behaviour of\nd(t)closely reflects the one of \u0001M(t).\nEffective Fermi energy and wavevector\nWe have defined \u0014Fand\u000fFas relevant length and energy scales.\nThese are evaluated by approximating the thin barrier at z= 0 as\na delta-like potential. This is reasonable considering that the bar-\nrier thickness of 2 \u0016m is about 70 times smaller than the typical\nThomas-Fermi radius of the cloud along the z-axis. Namely, we\napproximate the density distribution of a Fermi gas of Natoms\nconfined in a “half” of the harmonic trap bisected by the thin\nbarrier as half of the distribution of a Fermi gas of 2Natoms\nin the whole harmonic trap. The latter is evaluated using the\nfinite-temperature Fermi-Dirac distribution of an ideal Fermi gas\nnF(r;T=T F), given the measured T=TF, trap frequencies and atom\nnumber. From nF(r;T=T F)the local wavevector and Fermi en-\nergy are given by kF(r;T=T F) = (6\u00192nF(r;T=T F))1=3and\nEF(r;T=T F) = ~2k2\nF(r;T=T F)=(2mLi), respectively. The two\nparameters\u0014Fand\u000fFare derived by averaging kF(r;T=T F)and\nEF(r;T=T F)over a region as thick as the local interparticle spac-\ning(6\u00192)1=3=kF(x;y;0;T=T F)aroundz= 0 . Integration of\nnF(r;T=T F)within such region yields the total number Nintof\"\nand#fermions at the interface between the two spin domains (see\nSupplementary Information for details).7\nSum-rule approach for the spin-dipole mode frequency\nGiven a perturbation operator D, the moments of the strength\ndistribution function, or sum rules, are given by mk=P\nnjh0jDjnij2(En\u0000E0)k. In particular, the spin-dipole mode is\nexcited by the operator D=P\ni\"zi\u0000P\ni#zi, wherezis the lon-\ngitudinal coordinate and i\";#= 1;:::;N\";#. The frequency !SDof\nthe spin-dipole oscillation is estimated using the ratio (see Supple-\nmentary Information and Ref. 25):\n~2!2\nSD=m1\nm\u00001: (2)\nWithin the local density approximation (LDA) one can obtain (see\nSupplementary Information)\n!2\nSD=N\nmR\ndrz2\u001f(n); (3)\nwhere\u001f(n)is the zero-temperature magnetic susceptibility of a uni-\nform gas obtained at the density nfrom QMC calculations[20]. The\ndashed green line plotted in Fig. 2 is calculated through Eq. (3)\nby inserting the density profile n(r)of two fully overlapped spin\ncomponents, determined in LDA using the equilibrium condition\n\u0016(n)\u0000Vtrap=\u00160, where\u0016(n)is the chemical potential as a function\nof local density nfrom QMC calculations and \u00160is fixed by the nor-\nmalisation condition. The solid green line in Fig. 2 accounts instead\nfor a reduced overlap of 25 %around the trap centre, in closer anal-\nogy to the experimental condition. In this case, the integral in Eq. (3)\nover the outer spin-polarised regions contributes thus only with the\nspin susceptibility \u001f0of an ideal Fermi gas, yielding a reduced de-\nviation of the spin-dipole frequency from the bare trap frequency.\nTherefore, we expect the measured frequencies to be higher than the\nresults of Ref. 25, derived at full overlap. Consequently, the solid and\ndashed lines in Fig. 2 delimit a confidence region in which most ex-\nperimental data are found. Most importantly, the critical interaction\nstrength at which the abrupt change in !SDoccurs does not depend\non the initial overlap configuration. Moreover, the spin diffusion dy-\nnamics happens on a longer timescale ( &200ms), as displayed in\nthe Supplementary Information, compared to the spin-dipole period\n(\u001850\u0000100ms). Hence, while we cannot assume that the system\noscillates near an equilibrium configuration as in linear-response the-\nory, the (slow) timescale for diffusion is sufficiently separated from\nthe (faster) timescale for the spin oscillations.\nPolaron model for the domain-wall melting\nThe proposed modelling of the plateau data shown in Fig. 3 proceeds\nas follows. In the case of purely repulsive interaction, the ferromag-\nnetic state, if energetically allowed, would be indefinitely stable andthe miscibility of the two components would be prevented by the ex-\nistence of a domain wall. In particular, a #fermion at the interface\nwould need to pay a finite amount of energy \u001b > 0in order to ac-\ncess the other spin domain forming a repulsive polaron at energy E+.\nIn our metastable system, however, if a repulsive polaron is created,\nit can subsequently decay onto the lower branch with a rate \u0000, re-\nleasing an energy equal to the mismatch between the two branches,\nE+\u0000E\u0000. Hence, this two-step process will cause a net increase of\nenergy \u0001E=E+\u0000E\u0000\u0000\u001bat a rate \u0000. Importantly, the behaviour\nof\u0000,E+andE\u0000as a function of the interaction strength can be\nderived from recent non-perturbative theory approaches[12, 24]. We\nassume that at the beginning of the dynamics, the energy associated\nto the domain wall is given by \u001bN int,Nintbeing the total number of\nfermions within a slice around z= 0 of total thickness equal to one\ninterparticle spacing. The duration of the plateau \u001cpis then set by the\ncondition:\n\u001bN int= (E+\u0000E\u0000\u0000\u001b)\u001cp\u0000: (4)\nWe assume that \u001b=E+\u0000E+c, whereE+is the energy of a re-\npulsive polaron, while E+cis the energy of one free fermion at the\ninterface, left as a phenomenological parameter, and is independently\nadjusted for each T=TFherein investigated. From Eq. (S.19) we fit\nthe experimental data by optimising the only free parameter E+c(see\nalso Supplementary Information).\nDiffusion model for extracting the spin drag coefficient\nThe equation for the dynamics of the relative centre of mass d=\nz\"\u0000z#can be easily obtained from the Boltzmann equation and is\nwritten as (see Supplementary Information)\nd+ \u0000s_d+!2\nzd= 0; (5)\nwhere!zis the longitudinal trap frequency, and \u0000sthe spin drag co-\nefficient due to collisions[28, 29]. We obtain the experimental spin\ndrag coefficient by fitting the solution of Eq. (5) to the data, consid-\nering as initial condition d(0) =d0and_d(0) = 0 .\nIn Fig. 4 we compare the experimental results with a theoreti-\ncal prediction based on a T-matrix approximation for the scattering\ncross-section corrected by the available scattering states in order to\nhave a well defined scattering amplitude in the collisional integral\n(see Supplementary Information). The agreement is especially good\ndown to temperatures as low as T= 0:3TF. At lower tempera-\ntures, both the shape and the magnitude of the spin drag coefficient\nas a function of the interaction compare more poorly. This is ex-\npected since at very low temperature the T-matrix approximation is\nnot quantitatively correct, and moreover the gas may suffer some\nheating during the dynamics due to decay processes, making its tem-\nperature higher than the one measured at the start of the dynamics,\nwhich is used for the comparison with the theory model.\n[1] V ollhardt, D., Blumer, N. & Kollar, M. Metallic ferromagnetism –\nAn electronic correlation phenomenon . V ol. 580 of Lecture Notes in\nPhysics (Springer, 2001).\n[2] Brando, M., Belitz, D., Grosche, F. M. & Kirkpatrick, T. R. Metallic\nquantum ferromagnets. Rev. Mod. Phys. 88, 025006 (2016).\n[3] V ollhardt, D. & W ¨olfle, P. The superfluid phases of helium-3 (Taylor\nand Francis, 1990).\n[4] Silverstein, S. D. Criteria for ferromagnetism in dense neutron Fermi\nliquids-neutron stars. Phys. Rev. Lett. 23, 139 (1969).\n[5] Tatsumi, T. Ferromagnetism of quark liquid. Phys. Lett. B 489, 280\n(2000).\n[6] Stoner, E. Atomic moments in ferromagnetic metals and alloys with\nnon-ferromagnetic elements. Philos. Mag. 15, 1018 (1933).[7] Saxena, S. S. et al. Superconductivity on the border of itinerant-electron\nferromagnetism in UGe 2.Nature 406, 587 (2000).\n[8] Pfleiderer, C., Julian, S. R. & Lonzarich, G. G. Non-Fermi-liquid nature\nof the normal state of itinerant-electron ferromagnets. Nature 414, 427\n(2001).\n[9] Chin, C., Grimm, R., Julienne, P. S. & Tiesinga, E. Feshbach reso-\nnances in ultracold gases. Rev. Mod. Phys. 82, 1225–1286 (2010).\n[10] Shenoy, V . B. & Ho, T.-L. Nature and properties of a repulsive Fermi\ngas in the upper branch of the energy spectrum. Phys. Rev. Lett. 107,\n210401 (2011).\n[11] Kohstall, C. et al. Metastability and coherence of repulsive polarons in\na strongly interacting Fermi mixture. Nature 485, 615 (2012).8\n[12] Massignan, P., Zaccanti, M. & Bruun, G. M. Polarons, dressed\nmolecules, and itinerant ferromagnetism in ultracold Fermi gases. Rep.\nProg. Phys. 77, 034401 (2014).\n[13] Sanner, C. et al. Correlations and pair formation in a repulsively inter-\nacting Fermi gas. Phys. Rev. Lett. 108, 240404 (2012).\n[14] Lee, Y . et al. Compressibility of an ultracold Fermi gas with repulsive\ninteractions. Phys. Rev. A 85, 063615 (2012).\n[15] Pekker, D. et al. Competition between pairing and ferromagnetic insta-\nbilities in ultracold fermi gases near Feshbach resonances. Phys. Rev.\nLett. 106, 050402 (2011).\n[16] Duine, R. A. & MacDonald, A. H. Itinerant ferromagnetism in an ul-\ntracold atom Fermi gas. Phys. Rev. Lett. 95, 230403 (2005).\n[17] LeBlanc, L. J., Thywissen, J. H., Burkov, A. A. & Paramekanti, A. Re-\npulsive Fermi gas in a harmonic trap: ferromagnetism and spin textures.\nPhys. Rev. A 80, 013607 (2009).\n[18] Conduit, G. J., Green, A. G. & Simons, B. D. Inhomogeneous phase\nformation on the border of itinerant ferromagnetism. Phys. Rev. Lett.\n103, 207201 (2009).\n[19] Cui, X. & Zhai, H. Stability of a fully magnetized ferromagnetic\nstate in repulsively interacting ultracold Fermi gases. Phys. Rev. A 81,\n041602(R) (2010).\n[20] Pilati, S., Bertaina, G., Giorgini, S. & Troyer, M. Itinerant ferromag-\nnetism of a repulsive atomic Fermi gas: A quantum Monte Carlo study.\nPhys. Rev. Lett. 105, 030405 (2010).\n[21] Chang, S., Randeria, M. & Trivedi, N. Ferromagnetism in the upper\nbranch of the Feshbach resonance and the hard-sphere Fermi gas. Proc.\nNat. Acad. Sci. 108, 51 (2011).\n[22] Jo, G. et al. Itinerant ferromagnetism in a Fermi gas of ultracold atoms.\nScience 325, 1521 (2009).\n[23] Scazza, F. et al. Repulsive Fermi polarons in a resonant mixture of\nultracold6Li atoms. Phys. Rev. Lett. 118, 083602 (2017).\n[24] Schmidt, R. & Enss, T. Excitation spectra and RF response near\nthe polaron-to-molecule transition from the functional renormalization\ngroup. Phys. Rev. A 83, 063620 (2011).\n[25] Recati, A. & Stringari, S. Spin fluctuations, susceptibility and the dipole\noscillation of a nearly ferromagnetic Fermi gas. Phys. Rev. Lett. 106,\n080402 (2011).\n[26] Bienaim ´e, T. et al. Spin-dipole oscillation and polarizability of a binary\nBose-Einstein condensate near the miscible-immiscible phase transi-\ntion. Phys. Rev. A 94, 063652 (2016).\n[27] Sanner, C. et al. Speckle imaging of spin fluctuations in a strongly\ninteracting Fermi gas. Phys. Rev. Lett. 106, 010402 (2011).\n[28] Sommer, A., Ku, M., Roati, G. & Zwierlein, M. Universal spin trans-\nport in a strongly interacting Fermi gas. Nature 7342 , 201 (2011).\n[29] Enss, T. & Haussmann, R. Quantum mechanical limitations to spin\ndiffusion in the unitary Fermi gas. Phys. Rev. Lett. 109, 195303 (2012).\n[30] Bardon, A. B. et al. Transverse demagnetization dynamics of a unitary\nFermi gas. Science 344, 722 (2014).[31] Burchianti, A. et al. Efficient all-optical production of large6Li quan-\ntum gases using D1gray-molasses cooling. Phys. Rev. A 90, 043408\n(2014).\n[32] Valtolina, G. et al. Josephson effect in fermionic superfluids across the\nBEC-BCS crossover. Science 350, 1505 (2015).\n[33] Z ¨urn, G. et al. Precise characterization of6LiFeshbach reso-\nnances using trap-sideband-resolved RF spectroscopy of weakly bound\nmolecules. Phys. Rev. Lett. 110, 135301 (2013).\n[34] Duine, R. A., Polini, M., Stoof, H. T. C. & Vignale, G. Spin Drag in\nan ultracold Fermi gas on the verge of ferromagnetic instability. Phys.\nRev. Lett. 104, 220403 (2010).\n[35] Taylor, E., Zhang, S., Schneider, W., & Randeria, M. Colliding\nclouds of strongly interacting spin-polarized fermions. Phys. Rev. A\n84, 063622 (2011).\n[36] Nascimbene, S. et al. Fermi-Liquid behavior of the normal phase of\na strongly interacting gas of cold atoms. Phys. Rev. Lett. 106, 215303\n(2011).\n[37] Tajima, H., Hanai, R. & Ohashi, Y . Strong-coupling corrections to spin\nsusceptibility in the BCS-BEC-crossover regime of a superfluid Fermi\ngas. Phys. Rev. A 93, 013610 (2016).\n[38] Goulko, O., Chevy, F. & Lobo, C. Collision of two spin-polarized\nfermionic clouds. Phys. Rev. A 84, 051605 (2011).\n[39] Elliott, E., Joseph, J. A. & Thomas, J. E. Anomalous minimum in the\nshear viscosity of a Fermi gas. Phys. Rev. Lett. 113, 020406 (2014).\n[40] Trotzky, S. et al. Observation of the Leggett-Rice effect in a unitary\nFermi gas. Phys. Rev. Lett. 114, 015301 (2015).\n[41] Levinsen, J. & Parish, M. M. Strongly interacting two-dimensional\nFermi gases . V ol. 3, Chap. 1, pp. 1–75 of Annual Review of Cold Atoms\nand Molecules (World Scientific, 2015).\n[42] Pilati, S., Zintchenko, I. & Troyer, M. Ferromagnetism of a repulsive\natomic Fermi gas in an optical lattice: a quantum Monte Carlo study.\nPhys. Rev. Lett. 112, 015301 (2014).\n[43] Pilati, S. & Fratini, E. Ferromagnetism in a repulsive atomic Fermi gas\nwith correlated disorder. Phys. Rev. A 93, 051604(R) (2016).\n[44] He, L., Liu, X.-J., Huang, X.-G. & Hu, H. Stoner ferromagnetism of a\nstrongly interacting Fermi gas in the quasirepulsive regime. Phys. Rev.\nA93, 063629 (2016).\nAcknowledgments – We thank A. Morales and J. Seman for contributions\nin the early stage of the experiment, and G. Bertaina, G. M. Bruun, C. Di\nCastro, C. Fort, S. Giorgini, R. Grimm, W. Ketterle, P. Massignan, S. Pilati,\nR. Schmidt, W. Zwerger, M. Zwierlein and the LENS Quantum Gases group\nfor many stimulating discussions. We thank H. Tajima and Y . Ohashi for pro-\nviding us recent data of the lower branch spin susceptibility. This work was\nsupported under European Research Council grants no. 307032 QuFerm2D,\nand no. 637738 PoLiChroM. A.R. acknowledges support from the Alexan-\nder von Humboldt foundation. T.E. acknowledges the Physics Department,\nSapienza University of Rome, for hospitality, and the Humboldt foundation\nfor financial support during part of this work.\nAdditional information – Correspondence and requests for materials should\nbe addressed to M.Z. (e-mail: zaccanti@lens.unifi.it).1\nSUPPLEMENTARY INFORMATION\nExploring the ferromagnetic behaviour of a repulsive Fermi gas via spin dynamics\nG. Valtolina,1;2;3F. Scazza,1;2A. Amico,1;2A. Burchianti,1;2\nA. Recati,4;5T. Enss,6M. Inguscio,1;2M. Zaccanti1;2;\u0003and G. Roati1;2\n1INO-CNR, Via Nello Carrara 1, 50019 Sesto Fiorentino, Italy\n2LENS and Dipartimento di Fisica e Astronomia, Universit `a di Firenze, Via Nello Carrara 1, 50019 Sesto Fiorentino, Italy\n3Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy\n4INO-CNR BEC Center and Dipartimento di Fisica, Universit `a di Trento, 38123 Povo, Italy\n5Technische Universit ¨at M ¨unchen, James-Franck-Straße 1, 85748 Garching, Germany\n6Universit ¨at Heidelberg, Philosophenweg 19, 69120 Heidelberg, Germany\n\u0003E-mail: zaccanti@lens.unifi.it\nS.1. EXPERIMENTAL METHODS\nFermi gas preparation procedure\nOur procedure to create weakly interacting two-component Fermi mixtures of6Li atoms has been already described in detail\nelsewhere [31]. To realize highly degenerate samples at T=TF<0:2, a balanced mixture of atoms equally populating the lowest\nand third-to-lowest Zeeman states (hereafter denoted j1iandj3i, see Fig. S1a) is evaporated in a crossed optical dipole trap\n(ODT) at a bias magnetic field of 300 Gauss. At such a field the j1i\u0000j3imixture features a large [33], though off-resonant,\nscattering length value of about -900 a0which makes the forced evaporation process extremely efficient [31]. Here a0is the\nBohr’s radius. We typically end up with N1'N3'5\u0002104atoms per spin component, confined in a cigar-shaped harmonic\npotential with axial and radial trap frequencies \u0017z= 21:0(1) Hz and\u0017?= 265(5) Hz, respectively.\nFor the preparation of less degenerate samples, we either heat up in a controlled way the cloud by quickly turning off and\non the ODT for an excitation time up to 1 ms, or we perform evaporation at 300 Gauss with a j1i\u0000j2imixture. This features\nan interspecies scattering length of about -300 a0, which makes the evaporation process less efficient than for the j1i\u0000j3ispin\ncombination, leading to similar atom number and trap frequencies, though at higher temperature. The degeneracy parameter\nT=TFis determined by fitting the cloud density profiles in situ or after 5ms of time of flight to a finite temperature Fermi-Dirac\ndistribution.\nAt the end of evaporation in the j1i\u0000j3imixture, we transfer all atoms from the j3ito thej2istate via a resonant 80 \u0016s\nradio-frequency (RF) \u0019-pulse. To avoid detrimental finite-state interaction effects, that would limit the transfer efficiency, we\nperform the transfer at a bias field of 584.5 G, where the scattering lengths of the j1i\u0000j3iandj1i\u0000j2ispin mixtures are equal\nand non-resonant [33].\nSpin separation procedure\nThe procedure for creating two separate spin domains is described here and summarized in Fig. S1c. In order to spatially\nseparate thej1iandj2icomponents, we adiabatically lower the magnetic field down to \u00181G, where the two states are essen-\ntially non-interacting and possess equal but opposite magnetic moments (see Fig. S1b). Subsequently, we turn on a magnetic\nquadrupole gradient of about 1 G/cm along the weak trap axis through a 40 ms linear ramp, shifting the two spin clouds in\nopposite directions. After 180 ms, once the overlap between the two components is completely zeroed, we turn on through\na 40 ms linear ramp a strongly anisotropic 532 nm optical beam [32] centered at z= 0, and characterized by a short (long)\n1=e2waist ofwz=2.0(2)\u0016m (wx=840(30)\u0016m). The beam, blue-detuned with respect to the 671 nm lithium main transition,\nresults in a repulsive potential which bisects the cloud along the weak axis into two reservoirs. The typical maximum barrier\nheight after the 40 ms ramp is about 10 EF, a value that impedes any appreciable particle tunnelling between the two reservoirs\nover more than 2 s. Once the two spin clouds are separated and disconnected, we adiabatically turn off the magnetic gradient\nand increase the bias Feshbach field with a 500 ms linear ramp up to the target value. To prevent atom losses due to the mag-\nnetic field gradient, after the evaporation we turn on two additional elliptic plug beams at 532 nm with short (long) waist of\nwz= 50\u0016m (wy= 120\u0016m). The short waist is aligned along the weak trap axis and the plugs enter perpendicularly to this,\ncreating two effective repulsive walls that reduce as much as possible spilling of atoms due to the magnetic gradient. The plugs\nare turned on through a 200 ms linear ramp after evaporation and switched off through a 450 ms linear ramp when the Feshbach2\nFeshbac h\nB-fieldGradient\noff300 G584.5 Gtarget\nPlugsBarrier\noff off\n200 150 450 1030 5 450 5050305\ntime(ms)\n~2εF~10εF\n4030 1501 G\noffoff\n-180-170-160-150\n10 20 30 40100 200 300-400-2000200400\nMagn eticfield(G)Magn eticfield(G)Energylevels(MHz) Energylevels(MHz)a\nbc\nFIG. S1. a, Hyperfine and Zeeman levels structure of6Li in theF= 1=2andF= 3=2manifolds as a function of the magnetic field. b,\nEnergy dependence of the two lowest Zeeman states ( j1iandj2i) at low magnetic field. Below 5 G, the two states have magnetic moments\nwith nearly equal amplitude but opposite sign. c, The experimental sequence used for spin separation is sketched: the Feshbach field (light\nblue), the magnetic gradient (purple), the side plugs (dark green) and the central barrier (light green) are plotted as a function of time.\nfield is increased up to the final target value. The overall spin-separation procedure causes a decrease of about 25 %of the atom\npopulation, whereas no change of T=TFis measured. The degeneracy parameter after spin separation is estimated by a finite\ntemperature Fermi-Dirac distribution fit of the density distribution of spin polarized clouds recorded after a 5 ms time-of-flight\nexpansion at a bias field of 300 G.\nIt is important to stress that the final configuration is perfectly symmetric, both in total spin population and density distribution,\nand each component does not present any appreciable shape nor center-of-mass oscillation within the separated reservoirs.\nWe note here that a measurement of the oscillation frequency of a single spin component within one isolated reservoir yields\n\u0017R= 1:78(5)\u0017z. This is only about 10 %lower than the value 2 \u0017zexpected for the case of an infinitely thin barrier. This justifies\nthe approximation of the barrier potential as delta-like for modelling the cloud density, as discussed in detail in Section S.2.\nExcitation and measurement of spin response\nIn order to excite the spin-dipole mode at a fixed Feshbach field, we abruptly turn off the barrier potential from its maximum\nheight of about 10EFdown to zero, within less than one \u0016s. The two spin clouds, whose edges are initially separated by\na distance of about 5 \u0016m, start moving one towards the other at a small, though non-zero, relative velocity. From here on,\nwe monitor the dynamics of the relative distance between the cloud centers of mass. For each value of the hold time, in two\nindependent and successive experimental runs we acquire two in situ absorption images, each of which is resonant only with one\nspin state. For each value of interaction, temperature, and evolution time herein investigated, the relative distance between the\ntwo clouds is determined by averaging over at least five independent measurements.\n0 20 40 60 80 100 12080100120140\nHold time (ms)d(t)(µm)\nFIG. S2. Evolution of the center-of-mass distance between the two spin clouds, d(t) = (z\"\u0000z#)(t), at1=(\u0014Fa)'0:05andT=TF'0:12.\nThe dashed curve is obtained by fitting the data with the sum of an exponential decay and a damped sinusoidal oscillation.3\nThe center of mass ziof thei=\";#cloud is equivalently determined either via a Gaussian fit to the imaged density distribution\n(used for data displayed in Fig. 2a-c) or by directly evaluating zivia integration of the bare images. Once the spin dynamics is\nrecorded (see e.g. Fig. S2), we fit d(t) = (z\"\u0000z#)(t)to an exponential decay. By subtracting the fitted exponential decrease\nofd(t)due to diffusion from the data, we can isolate the spin-dipole mode, whose frequency is extracted from the data by\nmeans of a damped sinusoidal fitting function f(t) =Acos(2\u0019\u0017st)e\u0000t=\u001c. Similarly, direct integration of the images at each\nevolution time allows to determine the relative population of the i=\";#component in the left and right reservoirs, Mi=\n(Ni;L\u0000Ni;R)=(Ni;L+Ni;R), from which we obtain the total magnetization \u0001M= (M\"\u0000M#)=2. The behavior of d(t)\nclosely reflects the one of \u0001M(t).\nMeasurement of magnetization evolution and spin diffusion\nAs already mentioned in the main text, to investigate the spin diffusion of the two domains into each other, and in particular to\nreveal the dynamical arrest of spin diffusion, we employ a different protocol which does not excite the spin-dipole mode. For this\npurpose, we turn off the repulsive barrier slowly in a two-stage sequence. First, we lower the barrier from 10 EFdown to about\n\u00142EFthrough a linear 30 ms ramp (see Fig. S1c). The ramp speed and final barrier height are chosen to ensure an adiabatic re-\nadjustment of the density distributions of the two clouds in the barrier region, while preventing a detectable tunnelling of atoms\nacross the barrier and not inducing any shape excitations. In this way, the magnetization in each reservoir does not change, but\nthe relative distance between the edges of the two spin domains near z= 0 is drastically reduced, from 5 \u0016m down to about\n1\u0016m, a length-scale comparable with the mean interparticle spacing of our gas at the interface.\nIn a second stage, we ramp the intensity of the repulsive barrier down to zero. We have investigated different durations of the\nsecond ramp, spanning from 0 up to 30 ms. For ramps durations between 10 and 30 ms, we detect an appreciable flow of atoms\nacross the barrier region already during the ramp. In this case, once the barrier is off, the spin dynamics is well described by\na continuous, single exponential decay. For ramps shorter than 10 ms, in the strongly interacting regime, we observe a \u001cp>0\nwindow of vanishing spin diffusion. For ramp times below 5 ms, the duration of the plateaus is maximized and, for each target\nfield explored in this work, it does not show any dependence on the ramp speed. For the measurements presented in Fig. 3 and\n4 of the main text, the barrier is turned off in 5 ms. Examples of the short-time evolution of the cloud density profiles and of\nthe longitudinal magnetization M(z) = (n\"\u0000n#)(z)=(n\"+n#)(z)are displayed in Fig. S3. Consistently with the trend of the\ntotal magnetization \u0001Mshown in Fig. 3 of the main text, the evolution of the longitudinal magnetization at strong interactions\nexhibits a halt after a short evolution time (see Fig. S3b). After spin diffusion has been established at t >50ms, the evolution\nof the magnetization at each temperature and interaction strength is well described by Eq. (S.23) (see e.g. Fig. S4).\nt = 0 ms\nt = 10 ms\nt = 15 mst = 20 ms\nt = 70 ms\nt = 350 msb at = 0 ms\nt = 15 ms\nt = 50 ms\n[1 px = 1.04 µm]\nFIG. S3. a, Radially-integrated density profiles n\"(z)(blue) andn#(z)(red) of the two spin domains at different hold times after slowly\nremoving the optical barrier, with 1=(\u0014Fa)'0:05andT=TF'0:12.b, Longitudinal magnetization M(z)at different hold times for\n1=(\u0014Fa)'0:05(purple) and 1=(\u0014Fa)'0:9(green) atT=TF'0:12. It is visible here that M(z)remains substantially unchanged for\n1=(\u0014Fa)'0:05at hold times between 10 and 20 ms.4\n200 400 600 8000.050.10.51\nHold time (ms)∆M1/κF a = 0.05\n1/κF a = 1.35\n1/κF a = 2.401/κF a = -0.25\nFIG. S4. Spin diffusion dynamics at hold times t >50ms forT=TF= 0:12(2) . The evolution of the magnetization \u0001M(t)is plotted for\nvarious interaction strengths. Data points averaged over 4-5 indipendent experimental realizations (with error bars representing the s.e.m. of\nthe data) are shown together with fits to the solution of the simple kinetic model from which we extract \u0000S(see Eq. (S.23)).\nS.2. EFFECTIVE FERMI ENERGY AND FERMI WA VEVECTOR\nIn order to account for the inhomogeneity, the finite temperature, and the initial density of the two spin components in\ndisconnected reservoirs, in the main text we have defined \u0014Fand\u000fFas relevant length and energy scales. These are evaluated\nby approximating the thin barrier at z= 0 as a delta-like potential. As already discussed in Section S.1, this is a reasonable\nassumption, considering that the barrier thickness of 2 \u0016m is about 70 times smaller than the typical Thomas-Fermi radius of\nthe cloud along the zaxis. This is also confirmed by the measurement of the oscillation frequency within each reservoir, that is\nfound to be\u001810% lower than the value 2\u0017zexpected for the case of a delta-like potential. Namely, we approximate the density\ndistribution of a Fermi gas of Natoms confined in “half” a harmonic trap as half of the distribution of a Fermi gas of 2 Natoms\noccupying the whole dipole trap. The latter is evaluated using the finite-temperature Fermi-Dirac distribution of an ideal Fermi\ngasnF(r;T=TF), given the measured T=TF, trap frequencies and atom number (see Fig. S5a). The obtained profile excellently\nreproduces the two \"and#clouds, imaged in situ independently. From nF(r;T=TF)the local Fermi energy and wavevector are\ngiven respectively by kF(r;T=TF) = (6\u00192nF(r;T=TF))1=3andEF(r;T=TF) =~2kF(r;T=TF)2=(2mLi).\nThe two parameters \u0014Fand\u000fFare derived by averaging the local kF(r;T=TF)andEF(r;T=TF)over a region as thick as the\nlocal interparticle spacing (6\u00192)1=3=kF(x;y;0;T=TF)aroundz= 0(see Fig. S5b). Parallel to this, integration of nF(r;T=TF)\nwithin this region yields the total number Nintof\"and#fermions at the interface between the two spin domains. The definition\nof such an energy and length scale are inspired by a recent study on the spatial distribution of a fully ferromagnetic two-fermion\nmixture in a harmonic trap [45], which predicts that the domain wall at the interface between the two magnetic domains has a\ntotal thickness of one interparticle spacing.\n0050\n-5025\n-25\n5 -5 10 -10\nkF zkF y\n01\n00\n-100100\n-800 800 400 -400\nkF zkF y\narb. unitsa b\nFIG. S5. a, Model for total density distribution of the initial state: the optical barrier is approximated as infinitely thin, and the total profile of\nthe two spin components is given by the one of a Fermi gas of N\"+N#atoms at the given temperature, occupying the whole trap. b, The\nvalues of the effective Fermi energy \u000fFand wave vector \u0014Fare obtained by averaging over a slice (shown) around z= 0 of total thickness\nequal to one local interparticle spacing.5\nS.3. SUM-RULE APPROACH FOR THE PREDICTION OF THE SPIN-DIPOLE FREQUENCY\nThe spin-dipole mode is an oscillation of the relative position of the center of mass (COM) of each cloud of j\"iandj#i-atoms,\nbut not of their total COM position, i.e. a motion where atoms of one species move in the opposite direction with respect to the\natoms of the other species. For this reason, it is also referred to as an out-of-phase (spin) mode. As discussed in the main text,\nthe measurement of the spin-dipole mode frequency provides a spectroscopic tool for disclosing the spin susceptibility of the\ngas. An intuitive reasoning can clarify the relation between the mode frequency and the susceptibility, expressed in Eq. (1) of the\nmain text. A finite frequency of the spin-dipole mode is due to the presence of a restoring force when one tries to pull apart j\"i\nandj#i-particles in oppositedirections. With no interactions, the only restoring force is provided by the trapping potential and\nthe COM of each cloud oscillates at the trap frequency. A repulsive (attractive) interaction between the two atomic states leads to\na smaller (larger) energy cost to pull them apart with respect to the non-interacting case. This leads to a smaller (larger) restoring\nforce in the equation of motion, eventually resulting in a frequency smaller (larger) than the trap frequency. We indeed observe\na reduced(increased) frequency of the spin dipole mode when the system is prepared on the repulsive (attractive) branch (see\nFig. S7). For similar reasons, the spin susceptibility of a Fermi gas increases (decreases) if the interspecies particle interaction\nis repulsive (attractive). Therefore, a frequency decrease of the spin dipole corresponds to an increase of the spin susceptibility,\nand vice versa. The previous argument can be made more formal using a sum-rule approach [25, 46], which provides an upper\nbound of the spin-dipole mode frequency in terms of the system susceptibility. When the response is characterised by a single\ndelta-function (in the dynamic structure factor), the upper bound becomes an exact result. As emphasized in the main text, the\nmeasured spin-dipole oscillation frequency compares well with the sum-rule estimate supplemented with quantum Monte Carlo\n(QMC) data and employing a local density approximation [25].\nThe sum-rule approach is based on the knowledge of different momenta of the linear response of the system to the perturbation\nexciting the mode of interest. Given a perturbation operator D, the moments of the strength distribution function, or sum rules,\nare given by mk=P\nnjh0jDjnij2(En\u0000E0)k. An upper bound to the mode frequency is provided by the ratios of the sum\nrules. In particular, the spin-dipole mode is excited by the operator D=P\ni\"zi\u0000P\ni#zi, wherezis the longitudinal coordinate\nandi\";#= 1;:::;N\";#. We estimate the frequency of this mode using the ratio [25]:\n~2!2\nSD=m1\nm\u00001(S.1)\nThe reason is two-fold: (i) the inverse energy-weighted sum rule m\u00001is sensitive to low frequencies, which dominate since the\nstronger the repulsive interaction the softer the spin-dipole mode; (ii) m\u00001is related to the susceptibility of the system.\nThe so-called energy weighted sum rule m1is easily calculated in terms of a double commutator:\nm1=1\n2h0j[D;[H;D ]]j0i=N~2\n2m: (S.2)\nThe inverse energy-weighted sum rule is proportional to the spin dipole susceptibility. The latter can be determined by minimiz-\ning the total energy of the system in the presence of an external static coupling of the form \u0000\u0015D.\nSince our system is inhomogeneous, we can write the energy within the local density approximation (LDA):\nE=Z\ndr[\u000f(n\"(r);n#(r))\u0000\u0015z(n\"(r)\u0000n#(r))] (S.3)\nwhere\u000f(n\";n#)is the energy density of the uniform gas. By expanding \u000f(n\";n#)up to quadratic terms in n\"\u0000n#, minimization\nofEyields the result n\"\u0000n#=\u0015z\u001f(n)for the polarization induced by the external field. Here \u001f\u00001(n) =@2\u000f=@(n\"\u0000n#)2is\nthe zero-temperature inverse magnetic susceptibility of a uniform gas.\nThe calculation of the induced spin-dipole moment then yields m\u00001= 1=2R\ndrz2\u001f(n)and hence the result (Eq. (1) in the main\ntext)\n!2\nSD=N\"+N#\nmR\ndrz2\u001f(n)(S.4)\nfor the spin dipole frequency. Eq. (S.4) is exact within LDA and it shows that an increase of the magnetic susceptibility will\nresult in a decrease of !SD. If a ferromagnetic transition occurs, the spin-dipole mode frequency exhibits a minimum associated\nwith a diverging spin susceptibility within the central high-density region of the trap [25]. Once a ferromagnetic domain is\nformed, Eq. (S.4) is no longer valid and the dynamical behaviour of the system strongly depends on the geometry of the domain\nwall itself.\nFor an ideal Fermi gas trapped in a harmonic potential one gets the simple result !SD=!zwhere!zis the trap frequency\nalong thez-axis. For a balanced interacting Fermi gas on the repulsive branch, both for the energy density and the bulk magnetic6\nsusceptibility, we use the QMC results by Pilati et al. [20]. Both quantities can be written easily as an expansion in kFaabove\nthe mean-field results:\n\u000f\n\u000f0= 1 +kFa+C\u000f(kFa)2(S.5)\n\u001f0\n\u001f= 1\u00002\n\u0019kFa\u0000C\u001f(kFa)2; (S.6)\nwhere\u000f0and\u001f0are the chemical potential and the susceptibility of a spin- 1=2free Fermi gas, respectively. The constants\nC\u000f= 0:28andC\u001f= 0:62are obtained by fitting to the QMC results [20]. The previous expressions are suitable to determine\nan upper bound to the spin-dipole mode frequency for an unpolarized gas at equilibrium in the trap [25].\nProviding a rigorous theoretical description of the experimental results shown in Fig. 2 of the main text is challenging since\nthe gas has a time- and space-dependent local polarization. The dashed green line in Fig. 2 is calculated through Eq. (S.4) by\nassuming a cylindrically symmetric density profile n(r)with fully overlapped spin components, determined in LDA using the\nequilibrium condition \u0016(n)\u0000Vtrap=\u00160, where\u0016(n)is the chemical potential as a function of local density nfrom QMC\ncalculations and \u00160is fixed by the normalization condition. The solid green line in Fig. 2 accounts instead for a reduced overlap\nof 25 %around the trap center, in closer analogy to the experimental condition. In this case, the density profile used to compute\nEq. (S.4) is comprised of three longitudinal regions: two external non-interacting polarized tails and a central part where the\ntwo components fully overlap. For small oscillations, the overlap region can still be described through \u0016(n), i.e. the equation of\nstate of the repulsive branch. The size of the overlap region is chosen according to the typical experimental one. In particular,\na value of 25% denotes a ratio of 0.25 between the axial extension of the overlap region and the total axial size of the system.\nThe integral in Eq. (S.4) over the outer spin-polarized regions contributes thus only with the spin susceptibility \u001f0of an ideal\nFermi gas, yielding a reduced deviation of the spin-dipole frequency from the bare trap frequency. Therefore, we expect the\nmeasured frequencies to be higher than the results of Ref. 25, derived at full overlap. The solid and dashed lines in Fig. 2 delimit\na confidence region in which most experimental data are found. Most importantly, the critical interaction strength at which\nthe abrupt change in \u0017SDoccurs does not depend on the initial overlap configuration. Moreover, the spin diffusion dynamics\nhappens on a longer timescale ( &200ms), as displayed in Fig. S4, compared to the spin-dipole period ( \u001850\u0000100ms). Hence,\nwhile we cannot assume that the system oscillates near an equilibrium configuration as in linear-response theory, the (slow)\ntimescale for diffusion is sufficiently separated from the (faster) time scale of the spin oscillations.\nS.4. TEMPERATURE SHIFT OF THE CRITICAL INTERACTION STRENGTH\nWithin Landau-Fermi liquid theory the inverse susceptibility at T= 0can then be written as [47]\n\u001f\u00001(T= 0) =2\ng(eF)(1 +Fa\n0) (S.7)\nwhereFa\n0is thel= 0 antisymmetric (magnetic) Landau parameter, g(eF) =m\u0003kF=\u00192is the density of states with m\u0003=\nm(1 +Fs\n1=3)> m andFs\n1is thel= 1 symmetric (density) Landau parameter. To the first order in the interaction one has\nm\u0003=mandFa\n0=\u00002kFa=\u0019.\nAt finite temperature but T\u001cTFwe can consider the corrections to previous expression only due to free quasiparticles which\nis proportional to n, and therefore to T2. The interaction term depending on n2will contribute to higher order O(T4):\n\u001f\u00001(T) =2\ng(eF)\u0012\n1 +Fa\n0+\u00192\n12T2\nT2\nF\u0013\n(S.8)\nThe paramagnetic state becomes unstable when \u001f\u00001= 0, which at zero temperature can occur if there exists a critical value\n(kFa)csuch thatFa\n0=\u00001. At lowTwe can expand the Landau parameters around their values at the critical point (kFa)cat\nzero temperature\nFa\n0=Fa\n0;c+\u0012@Fa\n0\n@(kFa)\u0013\nc(kFa\u0000(kFa)c) (S.9)\nFs\n1=Fs\n1;c+\u0012@Fs\n1\n@(kFa)\u0013\nc(kFa\u0000(kFa)c): (S.10)\nTherefore we find that the critical temperature for the paramagnetic state to be unstable can be simply written as\n\u0012T\nTF\u0013\nc=2p\n3\n\u0019s\n\u0000\u0012@Fa\n0\n@(kFa)\u0013\ncp\nkFa\u0000(kFa)c (S.11)7\nFor instance using the expression for Fa\n0at the second order in kFa[16] one gets the first correction to the usual expression. In\nparticular (kFa)c'1:05and the critical temperature reads\n\u0012T\nTF\u0013\nc'23=2p\n3\n\u00193=2r\n1 + (kFa)c8\n3\u0019(1\u0000ln 2)p\nkFa\u0000(kFa)c (S.12)\nIn principle, application of the aforementioned power law as a fit to the experimental data in Fig. 3d is not justified: in fact,\n\u001cp>0is interpreted as the region of metastability of the fully ferromagnetic state, whereas Eq. (S.11) marks the boundary\nbetween paramagnetic and partially ferromagnetic phases. On the other hand, however, all QMC results show that the partially\nferromagnetic phase for a homogeneous system occupies a very narrow region of the phase diagram, and we expect that in\nour trapped system its presence is essentially washed out by the inhomogeneity of the clouds density distribution and finite\nexperimental resolution.\nS.5. PAIRING INSTABILITY AND MOLECULE FORMATION DURING SPIN DYNAMICS\nOne major issue that has hindered the study of strongly repulsive Fermi gases in previous experiments [13, 22] is represented\nby the pairing instability [15]. As discussed in the main text, ferromagnetic behaviour develops along the upper branch of\nthe many-body system: however, this state features an additional instability, represented by the tendency of the paramagnetic\nphase to turn via inelastic processes into a gas of pairs, which represents the true ground state of the balanced system at low\ntemperatures. At least for homonuclear mixtures and broad resonances [12, 13, 15], the pairing instability always overcomes the\nStoner’s one.\nThanks to our preparation scheme, that artificially initialises the system into a fully ferromagnetic configuration, we are able\nto contain the system tendency towards pairing, allowing for the investigation of the metastable upper branch. Furthermore, as\ndiscussed below in Section S.7, attractive polarons, rather than pairs, seem to be the preferential decay products in our system, at\nleast in the strongly interacting regime. Nonetheless, molecule formation has represented a major issue in previous experiments,\nand ruling out pairing effects for explaining the dynamics observed in our studies is fundamental for further supporting our\ninterpretation.\nIn previous experiments [13], the population of molecules and atoms has been identified via a rapid magnetic field ramp\ntechnique. After some evolution time at a target field close to the Feshbach resonance center, where molecules could be formed\nvia recombination processes, two successive absorption images were acquired. A first, taken at high field, allowed to monitor\nthe total population of atoms and molecules, since the molecule binding energy close to the Feshbach resonance is two to three\norders of magnitude smaller than the natural linewidth of the imaging transition. This allows to take a picture of the molecules\nwith the same imaging light employed for the atom, being the latter in the \"or#state. A second imaging pulse was taken after\na rapid sweep of the magnetic field to zero. The ramp converts weakly bound pairs into deeply bound molecules, which become\ntransparent to the atom imaging light. Hence, the second imaging selectively monitors the atom population, i.e. the population\nof the upper branch.\nOur setup does not allow to perform fast ( \u0018102G=\u0016s) ramps to low fields, hence we employed a different protocol to monitor\nthe presence of molecules in the system. This is based on acquiring, within a single experimental run, two subsequent absorption\nimages, by means of 4 \u0016s long pulses resonant with the \"and#states, respectively, and separated by 300 \u0016s. If no molecules\nare present, the effect of the first imaging pulse resonant with the \"state on#atoms is found to be negligible, since the \"\nimaging light is off-resonant to the #component. Furthermore, the short delay between the two pulses greatly limits the effect of\nheating of the#cloud associated to collisions with escaping \"imaged atoms. The effect of the first imaging pulse is completely\ndifferent if molecules are present. Since the \"\u0000# dimers are only weakly bound, the first imaging photon, resonant with the \"\noptical transition, dissociates the bound state into two atomic products [48], each of which symmetrically acquires a significant\nmomentum. The latter is associated to the density of states of the two outgoing atoms, to the binding energy of the dimer\n(negligible in this case), and to the photon momentum ~kL. The increase of the cloud size detected by the second imaging pulse\nis thus directly related to the amount of molecules in the system. We therefore monitor the increase of the radial width after the\nfirst pulse for different interaction strengths and different evolution times during the spin diffusion.\nA simple model allows to link the molecular fraction to the increase of the cloud width after the first pulse, following the\nexperimental protocol described above. In general, the size of a trapped cloud can be written as:\nhx2\n0i=2hUi\nm!2(S.13)\nWherehUiis the potential energy of one atom weighted over the density distribution of the cloud, the latter being eventually\nmodified by interaction effects. In the case of a pure gas of dimers, application of an imaging pulse resonant with the \"-\ncomponent causes the dimers to dissociate with a certain transfer of energy E1to the#-component. Since the photon energy is8\nalways larger than the binding one, we assume E1to be independent from the molecular binding, and hence independent from\n\u0014Fa. According to the same argument of Eq. (S.13), the width measured through the second pulse, following the first, can be\nwritten as:\nhx2\n1i=2hU+E1i\nm!2(S.14)\nIf the gas is a mixture of Nafree atoms and Nmmolecules, the mean size is set by:\nhx2i=Nahx2\n0i+Nmhx2\n1i\nNa+Nm(S.15)\nDefining the molecular fraction of the gas at time tduring the spin dynamics, fm\u0011fm(\u0014Fa;t) =Nm=N, we get:\nhx2i=2\nm!2hUi+2\nm!2hE1ifm=hx2\n0i+2\nm!2hE1ifm (S.16)\nWhen starting from a pure molecular sample, fm= 1, we would have:\n2\nm!2hE1i=hx2\n1mi\u0000hx2\n0mi (S.17)\nWe can therefore express the molecular fraction as:\nfm=hx2\n1i\u0000hx2\n0i\nhx2\n1mi\u0000hx2\n0mi(S.18)\nFor every investigated interaction strength, the denominator of Eq. (S.18) is experimentally determined by applying the double-\npulse imaging technique on a superfluid gas with the same interaction parameter \u0014Faand a temperature T=TF<0:1, ensuring\na molecular fraction fm'1on the BEC side of the resonance. The change of the density distribution when moving from\nthe unitary limit to the BEC one is accounted by renormalizing the measured radial width by the average density of the gas,\nevaluated with the conventional single-pulse absorption imaging.\nThe numerator of Eq. (S.18) is evaluated by measuring the second-pulse radial size of the cloud hx2\n1iat different evolution times\nduring the spin dynamics, initialized by the same experimental procedure discussed in Section S.1 and in the main text. hx2\n0i\nis in turn the size measured through the first imaging pulse at the corresponding times. Results of this analysis are reported in\nFig. S6, for various interaction strengths and different evolution times after abruptly removing the barrier, for a repulsive Fermi\ngas mixture initially prepared at T=TF= 0:12(2) .\nThe general trend for these measurements is interpreted as follows. In the weakly repulsive regime, the upper branch is very\nlong-lived, and despite rapid mixing of the two spin clouds only a small number of molecules is formed. Increasing interactions,\nMol.fraction50ms\n100ms\n1000ms\n00.10.20.30.40.50.60.7\n0 0.5 1 1.5 2 2.5\n1/κFa\nFIG. S6. Measured molecular fraction at T=TF= 0:12(2) for different evolution times after the barrier removal, obtained through the\ndouble-pulse imaging technique. The peak of molecule formation appears below the measured critical interactions, at 1=(\u0014Fa)'1:3.9\nthe decay rate of the upper branch monotonically increases [11, 12, 24, 49]: hence, despite an increase of the spin drag coefficient\n[28, 34], which tends to slow down the diffusion and reduces the spatial overlap of the \"and#clouds, the molecule formation\nbecomes more sizable, reaching a maximum near 1=(\u0014Fa)'1:3, in good agreement with Ref. 45. However, as one accesses the\nstrongly interacting regime where ferromagnetism is promoted, the molecule formation is again strongly reduced, highlighting\nthe tendency of the system to suppress the overlap between the two spin components. The trend at large \u0014Favalues, which\npersists also after long evolution times when the two clouds have come together, suggests that also at small values of local\npopulation imbalance the system may be a Fermi liquid state of attractive polarons, rather than a Bose gas of dimers. The Fermi\nliquid state might be favored by our way of initializing the system dynamics, as well as by the temperature increase associated\nwith the exothermic decay process from the upper to the lower branch. Importantly, for timescales below 100 ms, over which\nboth the magnetization plateau and the spin-dipole oscillations were measured, the observed heating is relatively small and the\nderived molecule fraction remains below 10 %for interaction strengths exceeding the critical value for the arrest of spin diffusion\nto occur. We therefore conclude that neither the behavior of the spin-dipole frequency nor the appearance of plateaus in the spin\ndiffusion can arise from dimer formation.\nThis conclusion is further supported by the measured temperature dependence of the critical repulsion strength (see Fig. 3d and\nEq. (S.11)). Such a trend is antithetic to the one that would result if the observed spin dynamics was governed by the molecular\npopulation within the overlap region. Assuming that such a molecular layer is in local chemical equilibrium with the atoms owing\nto fast thermalization, the molecular fraction is set by the Boltzmann factor Eb=(kBT)/1=(Ta2). Therefore, an increase of\nthe sample temperature would cause the same molecular fraction to form at smaller interaction strengths, corresponding to\nlarger molecule binding energies Eb, and leading to (T=TF)c/1=(\u0014Fa)2. This is qualitatively different from the observed\ntemperature dependence, displayed in Fig. 3d. While we cannot rule out a beneficial effect associated with the presence of\nmolecules at the interface, which may partially lower the tendency of the atoms to recombine into dimers, we can rule out that\ndimer formation lies at the origin of the observed dynamical behaviour.\nS.6. SPIN RESPONSE ON THE LOWER (ATTRACTIVE) BRANCH\nTo validate further our interpretation of the spin-dipole oscillation data presented in Fig. 2 and 3 of the main text, we have\nperformed additional spin response measurements, employing a system intentionally initialized in the lower energy branch. The\nresults are displayed in Fig. S7. The importance of such additional measurements is two-fold: on the one hand, we obtain novel\nexperimental hints on the trend of the lower-branch spin susceptibility via the same protocol employed for investigating the spin\nresponse in the upper branch; on the other hand, we can rule out that the trend of \u0017SD=\u0017zmeasured at strong coupling and\nshown in Fig. 2 of the main text could be ascribed to spin dynamics in the lower branch. From both experimental data and\ntheoretical calculations [27, 36, 37], the spin susceptibility of a Fermi gas on the lower branch of the Feshbach resonance is\n-3 -2 -1 0 10123\nνSD /νz\n-1/(κFa)\nFIG. S7. Spin-dipole frequency of an attractive Fermi gas at the BEC-BCS crossover. Red squares: \u0017SD=\u00170obtained after intentionally\ninitializing the system in the lower branch. Blue circles: spin-dipole frequency data shown in Fig. 2-3 of the main text, initializing the system\nto a repulsive Fermi gas. Green shaded region: theory prediction from Eq. (1) for a repulsive Fermi liquid, as shown in Fig. 2d of the main text.\nRed shaded region: theory prediction based on Eq. (S.4) for an attractive Fermi gas at T=TF= 0:2, obtained from the spin susceptibility and\nchemical potential given in Ref. [37]. The upper (lower) curve is obtained assuming full (20 %) spatial overlap between the two spin clouds.\nThe dashed red line denotes the theoretical expectation based on first-order perturbation theory [25].10\nknown to exhibit a monotonic decrease when passing from the BCS to the BEC side of the crossover. This is equally true for\nan attractive Fermi liquid and for a paired state, and for the latter both in the superfluid phase or in a pseudogap regime. In\nparticular, while the spin susceptibility of a weakly attractive system in the BCS limit tends to the value \u001f0of a non-interacting\nFermi gas, it strongly decreases when crossing the resonance and moving towards the BEC limit. Similarly, for all \u0014Favalues,\n\u001fis expected to monotonically decrease with decreasing temperature, especially if this drops below the critical temperature Tc\nfor superfluidity [37, 50]. Correspondingly, Eq. (1) of the main text (Eq. (S.4) here) predicts \u0017SD=\u0017z>1for an attractive Fermi\ngas at all temperatures and interaction strengths, and \u0017SDshould present a sharp increase when crossing the resonance from\na<0to thea>0(or once the temperature drops below Tc).\nThis supplementary characterization of the system dynamics is carried out by means of a slightly different experimental\nprotocol with respect to the one described in Section 1 and in the main text. We initially prepare a deeply degenerate system\natT=TF'0:12in the usual domain-wall configuration, at a bias field B0= 910 G, corresponding to \u0014Fa'\u00001. Here, the\nmeasured spin-dipole frequency, though still exceeding \u0017z, lies well below the value of \u0017SD= 1:70(4) obtained at\u0014Fa >1,\nand no arrest in the spin diffusion is ever observed. Consequently, we can assume that at this field, after the barrier removal, the\ngas immediately occupies the lower branch of the many-body system, as the upper one is neither well-defined nor energetically\naccessible.\nOnce the barrier is removed, we hold the field at B0for 20 ms, allowing a small mixing of the two spin clouds, and we\nsubsequently ramp it to a target value Btarget< B 0through a 25 ms ramp. We have checked that such a ramp is sufficiently\nadiabatic, so that the subsequent spin dynamics depends only weakly on the ramp duration. Once Btarget is reached, we\nmonitor the spin dynamics measuring the evolution of the relative distance between the centers-of-mass of the two spin clouds.\nThrough the same analysis procedure already described in the main text and Section 1 of the Supplementary Information, we\nextract\u0017SD=\u00170as a function the target interaction strength, as shown in Fig. S7 (red squares). It is evident how the spin-dipole\nfrequency measured after the adiabatic ramp from the BCS side, while being consistent with data shown in the main text for\n\u00001=\u0014Fa>0:5, presents a strong deviation further towards the BEC side. In this region, \u0017SDsharply increases to values above\n2\u00170already at unitarity, exceeding 3\u00170at1=\u0014Fa= 0:25. Further decreasing Btarget , hence\u0014Fa, no clear frequency can be\nextracted, owing to strong damping of the oscillations. Comparing the lower-branch data with the ones obtained by initializing\nthe system in the repulsive state (blue circles in Fig. S7) seems to indicate that the upper branch becomes completely unstable\ntowards pairing beyond unitarity at \u00001=\u0014Fa >0:5, such that the associated spin response is entirely compatible with the gas\noccupying the lower branch already at the beginning of the evolution.\nThe trend of our lower-branch data is well captured by Eq. (S.4), provided that we insert the spin susceptibility \u001fand chemical\npotential\u0016of a BEC-BCS crossover Fermi gas above the critical temperature for superfluidity. For our theory prediction (shown\nin Fig. S7 as the red shaded region), we have used recent data obtained within an extended T-matrix approximation and including\npairing fluctuations [37], which match both QMC results at unitarity [50] and experimental measurements of the lower-branch\nspin susceptibility [27].\nThe validity of the theoretical treatment based on Eq. (S.4) is therefore successfully confirmed in the well-established case\nof a BEC-BCS crossover attractive Fermi gas and our Supplementary Data of \u0017SDmatch previous experimental measurements\nof\u001fby Sanner et al. [27]. Finally, the reasonable agreement we find between our theoretical prediction and experimental data\nsuggests that our system is in a pseudo-gap regime of pre-formed pairs, rather than in an attractive Fermi liquid phase. For\nthe latter case, in fact, both experimental investigation and QMC studies at zero temperature [36] reported a value for the spin\nsusceptibility at unitarity that would lead to a spin-dipole frequency significantly lower than the one we measure. Such an\nintriguing aspect definitely will be the subject of future investigations.\nS.7. THEORETICAL MODEL FOR THE DOMAIN-WALL MELTING\nIn this Section, we describe the phenomenological model developed to explain the finite duration of the spin diffusion plateaus,\nas a function of interaction and temperature. Such a model is based on the precise knowledge of the spectral properties and\nlifetime of#impurities embedded in a Fermi sea of \"particles [24]. As discussed in the main text, the spectral function of\nsuch system is characterized by an upper and a lower branch. Depending on the interaction strength, associated to the lower\nbranch there exist in the extremely imbalanced case two kinds of Landau’s quasi-particles, coined attractive polarons and dressed\nmolecules [12]. For interaction parameters 1=(\u0014Fa)>0:9(1=(\u0014Fa)<0:9) dressed molecules (attractive polarons) represent\nthe absolute ground state of the many-body system. The upper branch, in turn, features a third kind of quasi-particle, termed\nrepulsive polaron, whose existence has been experimentally demonstrated both in a three-dimensional mass-imbalanced Fermi\nmixture of40K and6Li atoms [11], and also in a two-dimensional Fermi mixture of40K atoms [51]. While the lower branch is\nassociated to a net attractive interaction between the impurity and the particles of the medium, the upper one requires repulsion\nbetween the two species in order to develop. For increasing repulsive interaction ( 1=(\u0014Fa)!0+), the energy of the repulsive\npolaron progressively increases. However, parallel to this, the repulsive polaron acquires a progressively shorter lifetime, set by\nthe tendency of the system to decay onto the lower-lying states of the attractive branch. The decay rate \u0000associated with such11\ninelastic processes, together with the energies E+andE\u0000of repulsive and attractive polarons, has been determined through\nnon-perturbative theory approaches [12, 24].\nAlthough our system is a 50-50 balanced mixture of \"and#fermions, the results obtained in the impurity limit are extremely\nrelevant for understanding the existence of a plateau of zero diffusivity in the spin dynamics. Since our studies start by preparing\na phase-separated state, rather than a mixed paramagnetic phase, the initial mixing processes at the interface between the two spin\ndomains can be viewed as events in which fermions of one kind enter in the Fermi sea of the other component, and vice-versa.\nThose can be described in terms of dressed quasiparticle properties derived in the impurity limit.\nOur proposed explanation of the conductance plateau shown in Fig. 3 of the paper proceeds along the following line of thought.\nIn the case of purely repulsive interactions, the ferromagnetic state, if energetically allowed, would be indefinitely stable, and\nin a system with separately fixed spin populations would correspond to a phase-separated state. In fact, the miscibility of the\ntwo components would be prevented by the existence of a domain wall: namely, a #fermion at the interface would need to\npay a finite amount of energy \u001b > 0in order to access the other spin domain, forming a repulsive polaron. In our system,\nhowever, if a repulsive polaron is created, it can subsequently decay onto the lower branch with a rate \u0000, releasing an energy\nequal to the mismatch between the two branches, E+\u0000E\u0000. Hence, this two-step process will cause a net increase of energy\n\u0001E=E+\u0000E\u0000\u0000\u001bat a rate \u0000. We assume that at the beginning of the dynamics, the energy associated to the domain wall\nis given by \u001b\u0002Nint,Nintbeing the total number of fermions within a slice around z= 0 of total thickness equal to one\ninterparticle spacing (see Section S.2). The duration of the plateau \u001cpis then set by the condition:\n\u001bNint= (E+\u0000E\u0000\u0000\u001b)\u001cp\u0000 (S.19)\nWe write\u001b=E+\u0000E+c, whereE+cdenotes the energy of one free fermion at the interface. In homogeneous systems and\nin the zero-momentum impurity limit [12], E+c=EF. In our inhomogeneous sample at finite temperature, and in the vicinity\nof the interface between the two spin clouds, we introduce E+cas a phenomenological fitting parameter. Consequently, from\nEq. (S.19) we obtain the following expression for \u001cp:\n\u001cp=Nint(E+\u0000E+c)\n\u0000(E+c\u0000E\u0000)1\n2\u0019\u0017F(S.20)\nwhereh\u0017F=\u000fF. To compare the prediction of Eq. (S.20) to our data we employ the values of E+,E\u0000and\u0000given as a function\nof interactions by Schmidt et al. [24], and the values of \u0014F,\u000fFandNintobtained after radial averaging over finite temperature\ndensity profiles as described in Section S.2.\nFor each temperature regime herein investigated, the only free parameter of the model, E+c, is fixed by optimizing the\nagreement between experimental data and the prediction of Eq. (S.20). E+cnon-linearly increases upon increasing temperature:\nthis is expected since E+croughly corresponds to the energy of the repulsive polaron branch at the \u0014Favalue at which a first\nnon-zero\u001cpis detected. It is worth stressing how such a simple theory model, which accounts for finite temperature effects only\nviaE+cand a renormalization of \u0014F,\u000fFandNint, is able to provide a quantitative description of the experimentally observed\ntrend.\nThe model prediction stops at the unitary point and does not extend on the a <0side of the Feshbach resonance, because\nthe theory becomes unreliable in this region, the decay rate of the upper branch reaching the order of the Fermi energy, and\nhence making the repulsive polaron an ill-defined quasiparticle [24]. As a matter of fact, we are at present unable to provide a\ndescription for such a feature (and for its disappearance) on the BCS side of the resonance, which we cannot exclude to arise\nfrom a combination of collisional and many-body effects. Certainly, the measured \u001cpand\u0017SDat low temperatures are consistent\nwith the system temporarily accessing the upper branch even for \u0014Fa<0[15]. This extremely exotic many-body state, which\nis a three-dimensional analogue of the super-Tonks regime in one dimension [52], is thus far poorly explored and definitively\ndeserves deeper theoretical and experimental investigation.\nS.8. SPIN DRAG COEFFICIENT AND THEORY MODEL FOR THE DIFFUSION OF AN IMPURITY IN A FERMI SEA\nThe equation for the diffusion dynamics of the relative center of mass d(t) =z\"\u0000z#can be easily obtained starting from the\nBoltzmann equation and using the method of the averages [53]. One obtains the following coupled equations:\n@t(z\"\u0000z#)\u0000(v\"\u0000v#) = 0\n@t(v\"\u0000v#) +!2\nz(z\"\u0000z#) = (@t(v\"\u0000v#))coll; (S.21)\nwhere!zis the trapping frequency in the direction of the motion, v\"(v#) is the average velocity of the \"(#) spin component\nand(@t(v\"\u0000v#))collis the collisional term. The effect of the medium is included in this approach only in the collisional\nterm. Assuming that the distribution function changes in time only through the change of velocity, the collisional term is simply\nproportional to the relative velocity and can be written as\n(@t(v\"\u0000v#))coll=\u0000\u0000S(v\"\u0000v#); (S.22)12\nwhere \u0000Sis the so-called spin drag coefficient due to collisions [28, 29]. Therefore, substituting Eq. (S.22) in Eq. (S.21) the\nequation of motion for dis simply given by the one of a damped harmonic oscillator:\nd+ \u0000S_d+!2\nzd= 0 (S.23)\nWe obtain the experimental spin drag coefficient by fitting the solution of Eq. (S.23) to the data, with initial conditions d(0) =d0\nand_d(0) = 0 , or equivalently \u0001M(0) = \u0001M0and_\u0001M(0) = 0 (see Fig. S4).\nA main theoretical task is to determine the spin drag coefficient. For #impurities moving through a fully polarized noninter-\nacting\"Fermi sea, the drag rate can be computed as\n~\u0000S=2\u0019\nkBTn#Zd3p1\n(2\u0019~)3d3p2\n(2\u0019~)3d3p3\n(2\u0019~)3\u000e(\"1+\"2\u0000\"3\u0000\"4)\n\u0002(4\u0019~2)2\nm2d\u001b\nd\np1j(v1j\u0000v3j)n1#n2\"(1\u0000n3#)(1\u0000n4\"):(S.24)\nThis collision integral describes the scattering of an impurity atom with momentum ~ p1and a medium atom with momentum ~ p2\ninto outgoing states ~ p3and~ p4=~ p1+~ p2\u0000~ p3, conserving the total momentum and the total kinetic energy \"1+\"2of both\nparticles. Each particle has kinetic energy \"~ p=p2=2mand massm. The drag rate is proportional to the change in impurity\nvelocity,v1j\u0000v3j, wherejdenotes the spatial component in the direction of the initial impurity velocity. The scattering process\noccurs with probability n1#n2\"that the initial states are occupied, and probability (1\u0000n3#)(1\u0000n4\")that the final states are\nunoccupied, where n~ p\u001b= [exp(\f(\"~ p\u001b\u0000\u0016\u001b)) + 1]\u00001is the Fermi-Dirac distribution at chemical potential \u0016\u001b. The drag rate\nEq. (S.24) is derived from spin diffusion theory for a polarized Fermi gas [54–56] in the limit of vanishing minority density n#\n(the integral over the impurity distribution n1#cancels the factor n#in the denominator of Eq. (S.24) to yield a finite drag rate).\nEquivalently, Eq. (S.24) is obtained from the impurity drag rate [34, 57] in the limit of vanishing impurity velocity.\nIn order to compute the cross section d\u001b=d \nfor a dilute Fermi gas we use the T matrix in the ladder approximation [58]. In\nan ultracold Fermi gas the bare interaction is a s-wave contact interaction between unequal spins, thus also the T matrix only has\nans-wave component `= 0. One can then express the differential cross section\nd\u001b\nd\n=jf`=0(~ q;!)j2(S.25)\nin terms of the s-wave scattering amplitude f`=0(~ q;!)of two incoming particles with total momentum ~ q=~ p1+~ p2and total\nkinetic energy ~!=\"~ p1\"+\"~ p2#. The scattering amplitude, in turn, is given in terms of the T matrix T`as [58]\nf`=0(~ q;!) =\u0000mQ\n4\u0019~2T`=0(~ q;!): (S.26)\nFor two particles scattering in vacuum ( Q= 1), the two-body T matrix reads\nT(0)\n`=0(~ q;!) =4\u0019~2a=m\n1 +iak(S.27)\nwhereadenotes the s-wave scattering length, and ~~k= (~ p1\u0000~ p2)=2is the relative momentum of incoming particles. This\nresults in a vacuum scattering amplitude f(0)\n`=0=\u0000a=(1 +iak)and vacuum scattering cross section d\u001b(0)=d\n =jf(0)\n`=0j2=\na2=(1 +a2k2). At weak coupling jkF\"aj\u001c 1, the drag rate is proportional to the scattering cross section, \u0000S/a2. Note\nthat with the vacuum cross section, the drag rate depends only on the modulus jajof the scattering length and is symmetric\ninaaround unitarity ( a\u00001= 0) in the BCS-BEC crossover. The experimental data for the drag rate, however, exhibit a small\nasymmetry in a. A similar asymmetry in transport coefficients has been observed in the shear viscosity [39] and transverse spin\ndiffusion [40].\nIn order to explain the asymmetry in the drag rate it is necessary to include medium scattering, where the Fermi sea is Pauli\nblocked for intermediate states, and which entails a tendency toward molecule formation on the BEC side [59–62]. Medium\nscattering is described by the many-body T matrix T`,\nT\u00001\n`=0(~ q;!) =T(0)\u00001\n`=0(~ q;!) +Zd3p\n(2\u0019~)3n~ p#+n~ q\u0000~ p\"\n!\u0000\"~ p#\u0000\"~ q\u0000~ p\"+i0: (S.28)\nThe medium T matrix includes the effect of quantum degeneracy, which leads to a large increase in the partial-wave scattering\namplitudes (S.26) and would by itself violate the unitarity bound\njkf`j\u00141; (S.29)13\nwhich is the prerequisite for expressing the scattering amplitude kf`=ei\u000e`sin\u000e`in terms of real phase shifts \u000e`(~ q;!). The\ndefinition of the scattering amplitudes (S.26) in the presence of the medium therefore includes a phenomenological Qfactor\nwhich accounts for the fact that only unoccupied states are available for outgoing waves [58] (Pauli blocking),\nQ=Z1\n\u00001dcos\u0012\n2\u0010\n1\u0000n~ q=2+~~k#\u0000n~ q=2\u0000~~k\"\u0011\n: (S.30)\nThe integral averages over the angle \u0012between total momentum ~ qand relative momentum ~~k, while the modulus of ~kis fixed\nby the condition ~!=\"~ q=2+~~k#+\"~ q=2\u0000~~k\"=~2k2=m+q2=4m. In vacuum Q= 1, and also at high temperatures T\u001dTF\"\none hasQ\u00191. However, in a degenerate Fermi gas Q < 1lowers the scattering amplitude sufficiently to always satisfy the\nunitarity bound (S.29) also in the case of medium scattering.\nThe drag rate Eq. (S.24) is known analytically in the whole BCS-BEC crossover in the high-temperature limit T\u001dTF\"\nwhere the majority Fermi gas is non-degenerate [56],\n\u0010~\u0000S\nEF\"\u0011\nhightemp=16p\n2\n9\u00193=2\u0010TF\"\nT\u00111=2\u0002\n1\u0000x\u0000x2exEi(\u0000x)\u0003\nx=~2=(ma2kBT)(S.31)\nwhere Ei denotes the exponential integral. In this regime, the medium factor Q= 1. Conversely, medium scattering becomes\nimportant in a degenerate Fermi gas ( T.TF\"). In general, the collision term in Eq. (S.24) can be reduced to a three-dimensional\nintegral which is readily evaluated numerically. The resulting drag rate exhibits a maximum slightly on the BEC side of the\nresonance, in agreement with our experimental results. As the temperature is increased, the maximum shifts toward unitarity\nin agreement with the high-temperature expression (S.31) which is symmetric in a. At very low temperatures the agreement\nbetween this theory prediction and the experimental data is poorer, since there the T-matrix approximation is known to be\nquantitatively incorrect. Moreover the gas may suffer some heating during the dynamics due to decay processes, making its\ntemperature higher than the one measured at the start of the dynamics, which is used for the theory comparison.\nS.9. BREATHING MODE CHARACTERIZATION AND COLLISIONAL EFFECTS\nIn this Section we elaborate on the role of collisional effects, which could in principle dominate the system collective dynamics\nleading to misinterpretation of the observed spin-dipole mode behaviour. In strongly interacting many-body systems, collective\nmodes can be crucially affected by collisional hydrodynamics [60]. Within a purely collisional framework, the amplitude of\nspin-dipole oscillations is governed by the differential equation (S.23), which we use to analyze the long-time spin diffusion\ndynamics. Neglecting in-medium corrections, the collisional approach is based on the two-body scattering cross-section \u001b(k) =\n4\u0019a2=(1 + (ka)2)for relative momentum k. This leads to predictions independent of the sign of aand irrespective of whether\nthe system is prepared on the upper or lower branch, in clear contrast with our observations summarized in Fig. S7. This would\nnot change if one takes in-medium corrections into account (see Section S.8): at low temperatures these would lead only to a\nsmall asymmetry in the collisional integral [60], letting for instance many transport coefficients reach their maxima or minima\nslightly on the BEC side of resonance (see e.g. Fig 4 of the main text).\nLet us now consider the upper branch data for 0< \u0014Fa <1: the collisional hydrodynamics prediction (S.23) for the spin-\ndipole frequency might appear qualitatively compatible with the observed softening, given that \u0000Sincreases with interactions\n[38] (see Fig. 4 of the main text). As long as \u0000S<2!z, Eq. (S.23) would indeed give a spin-dipole frequency !SD=\n(!2\nz\u0000\u00002\nS=4)1=2. However, the damping rates \u0000Sexperimentally extracted by fitting the long-time evolution with the full\nsolution of Eq. (S.23) greatly exceed !z, yielding imaginary frequencies. On the other hand, we can consider Eq. (S.23) as a\nphenomenological model for the spin-dipole oscillations. In this case, the damping rate is the value 1=\u001cas extracted from the\nfitting procedure described in Section S.1, and therefore the measured spin-dipole frequency should be !SD= (!2\nz\u00001=\u001c2)1=2.\nOur fitted parameters do not satisfy the previous relation and more generally a collisional picture. Indeed the fitted \u001cvalues (i)\nwould give a very small correction to the bare frequency for any interaction strength and temperature and (ii) do not decrease by\nincreasing the value of \u0014Fa. Against a simple collisional interpretation speaks also the larger \u0017SDfound at higher temperature\nfor fixed\u0014Fa <1(purple squares in Fig. 2d). In the collisional dynamics framework the oscillation frequency should instead\ndecrease with temperature in the degenerate regime, where the damping rate increases because Pauli blocking becomes gradually\nless effective, as shown in Fig. 4.\nBreathing dynamics\nBesides exciting the spin-dipole mode, our experimental protocol also excites weakly the axial breathing mode, which we\nseparately characterize by studying the evolution of the cloud axial width \u001bz. Similarly to what we have done for the COM\nseparation dynamics, we isolate the breathing oscillation by subtracting from \u001bz(t)an overall exponential drift and fitting\nthe residual modulation \u0001\u001bz(t)with a damped sinusoidal function, as in Fig. S8a. The resulting frequency and damping14\nκF a0.1 1 101.21.41.61.82.0\nνB / νz\n√12/5\nκF a0.1 1 10Damping time (ms)\n50100150250\n200\n0\n-4-2024∆σz (µm)\n0 20 40 60 80 100\n0 20 40 60 80 100-202\nTime (ms)∆σz (µm)\n-505∆σz (µm)\n0 20 40 60 80 100\n-20-1001020∆σz (µm)\n0 20 40 60 80 100κF a = 0.04\nκF a = 0.29\nκF a = 0.77\nκF a ~ 50a b\nc\nFIG. S8. Characterization of the axial breathing mode. a, Evolution of the axial width \u0001\u001bz(t)for different interaction strengths at T=TF=\n0:12(2) after a sudden barrier switch-off. b, Normalized breathing frequency as a function of \u0014Fa. The breathing mode frequency \u0017B=\u0017z\ndecreases from the collisionless value 2, reaching a value betweenp\n12=5and2already at\u0014Fa'0:2.c, Damping time of the breathing\nmode versus \u0014Fa. The transition from a collisionless to a collisional regime is additionally signalled by a minimum of the damping time at\n\u0014Fa'0:2.\ntime are displayed in Fig. S8b-c. For weak interactions, the axial breathing mode oscillates at twice the axial trap frequency\n\u0017z, as expected in the collisionless regime. Increasing the interaction strength, we observe a decrease of the breathing mode\nfrequency\u0017B, which for\u0014Fa&0:2reaches an approximately constant value slightly larger than the hydrodynamic expectation\n\u0017B=p\n12=5\u0017z'1:55\u0017z[63] (see Fig. S8b).\nIn contrast to Ref. [28], we find the breathing mode amplitude to be substantially smaller than the one of the out-of-phase\nCOM oscillation at all interactions, owing to the small initial relative momentum imparted to the clouds by our experimental\nprotocol. For this reason, the spin-dipole mode is largely decoupled from the breathing one as long as the two clouds do not\nstart bouncing off each other, as if they became impenetrable. In this latter regime, observed at strong repulsion, the two modes\nare coupled and possibly lock to a common frequency within the experimental uncertainty. Such a frequency \u0017SD'1:70(4)\u0017z\nis compatible to \u0017R= 1:78(5)\u0017zthat we measured for a spin-polarized cloud oscillating in “half” the trap when the optical\nbarrier is left on; it is also similar to the experimental value 1:63\u0017z[28], and especially to the theory prediction 1:78\u0017z[35]. The\nvalue is also close to the frequencyp\n12=5\u0017zof the breathing mode in the hydrodynamic regime. In Ref. 38, where a purely\ncollisional approach is developed for two colliding clouds, it is found that the breathing mode remains essentially collisionless\nas long as the two clouds can diffuse into each other rapidly. Conversely, when the diffusion is slow, the breathing and the\nspin-dipole modes become coupled and the two clouds are predicted to enter into a bouncing regime. In this regime, both modes\nare locked to the same frequency which coincides at low temperature with the hydrodynamic predictionp\n12=5\u0017z[38]. In our\nexperiment, the bouncing is observed only deep in the collisional regime, and therefore our measurements of the dipole and the\nbreathing modes do not fit even qualitatively to a purely collisional approach. We further note that in the bouncing regime the\ntwo modes start being locked irrespective of the mechanism that makes the two clouds almost immiscible (see also Ref. 35).\nFurthermore, ascribing the bouncing dynamics to a collisional origin is completely inconsistent with the distinct behaviour of\nthe spin-dipole frequency measured on the lower branch, where \u0017SDis larger (smaller) than 1:55\u0017zat\u0014Fa'2(\u0014Fa'\u00001).15\nThe characterization of the breathing mode, together with the measured lower-branch spin-dipole frequency (see Fig. S7), lead\nus therefore to attribute the sudden jump of \u0017SDin Fig. 2d to a genuine system immiscibility, ruling out a dynamical origin.\n[1] V ollhardt, D., Blumer, N. & Kollar, M. Metallic ferromagnetism – An electronic correlation phenomenon . V ol. 580 of Lecture Notes in\nPhysics (Springer, 2001).\n[2] Brando, M., Belitz, D., Grosche, F. M. & Kirkpatrick, T. R. Metallic quantum ferromagnets. Rev. Mod. Phys. 88, 025006 (2016).\n[3] V ollhardt, D. & W ¨olfle, P. The superfluid phases of helium-3 (Taylor and Francis, 1990).\n[4] Silverstein, S. D. Criteria for ferromagnetism in dense neutron Fermi liquids-neutron stars. Phys. Rev. Lett. 23, 139 (1969).\n[5] Tatsumi, T. Ferromagnetism of quark liquid. Phys. Lett. B 489, 280 (2000).\n[6] Stoner, E. Atomic moments in ferromagnetic metals and alloys with non-ferromagnetic elements. Philos. Mag. 15, 1018 (1933).\n[7] Saxena, S. S. et al. Superconductivity on the border of itinerant-electron ferromagnetism in UGe 2.Nature 406, 587 (2000).\n[8] Pfleiderer, C., Julian, S. R. & Lonzarich, G. G. Non-Fermi-liquid nature of the normal state of itinerant-electron ferromagnets. Nature\n414, 427 (2001).\n[9] Chin, C., Grimm, R., Julienne, P. S. & Tiesinga, E. Feshbach resonances in ultracold gases. Rev. Mod. Phys. 82, 1225–1286 (2010).\n[10] Shenoy, V . B. & Ho, T.-L. Nature and properties of a repulsive Fermi gas in the upper branch of the energy spectrum. Phys. Rev. Lett.\n107, 210401 (2011).\n[11] Kohstall, C. et al. Metastability and coherence of repulsive polarons in a strongly interacting Fermi mixture. Nature 485, 615 (2012).\n[12] Massignan, P., Zaccanti, M. & Bruun, G. M. Polarons, dressed molecules, and itinerant ferromagnetism in ultracold Fermi gases. Rep.\nProg. Phys. 77, 034401 (2014).\n[13] Sanner, C. et al. Correlations and pair formation in a repulsively interacting Fermi gas. Phys. Rev. Lett. 108, 240404 (2012).\n[14] Lee, Y . et al. Compressibility of an ultracold Fermi gas with repulsive interactions. Phys. Rev. A 85, 063615 (2012).\n[15] Pekker, D. et al. Competition between pairing and ferromagnetic instabilities in ultracold fermi gases near Feshbach resonances. Phys.\nRev. Lett. 106, 050402 (2011).\n[16] Duine, R. A. & MacDonald, A. H. Itinerant ferromagnetism in an ultracold atom Fermi gas. Phys. Rev. Lett. 95, 230403 (2005).\n[17] LeBlanc, L. J., Thywissen, J. H., Burkov, A. A. & Paramekanti, A. Repulsive Fermi gas in a harmonic trap: ferromagnetism and spin\ntextures. Phys. Rev. A 80, 013607 (2009).\n[18] Conduit, G. J., Green, A. G. & Simons, B. D. Inhomogeneous phase formation on the border of itinerant ferromagnetism. Phys. Rev.\nLett. 103, 207201 (2009).\n[19] Cui, X. & Zhai, H. Stability of a fully magnetized ferromagnetic state in repulsively interacting ultracold Fermi gases. Phys. Rev. A 81,\n041602(R) (2010).\n[20] Pilati, S., Bertaina, G., Giorgini, S. & Troyer, M. Itinerant ferromagnetism of a repulsive atomic Fermi gas: A quantum Monte Carlo\nstudy. Phys. Rev. Lett. 105, 030405 (2010).\n[21] Chang, S., Randeria, M. & Trivedi, N. Ferromagnetism in the upper branch of the Feshbach resonance and the hard-sphere Fermi gas.\nProc. Nat. Acad. Sci. 108, 51 (2011).\n[22] Jo, G. et al. Itinerant ferromagnetism in a Fermi gas of ultracold atoms. Science 325, 1521 (2009).\n[23] Scazza, F. et al. Repulsive Fermi polarons in a resonant mixture of ultracold6Li atoms. Phys. Rev. Lett. 118, 083602 (2017).\n[24] Schmidt, R. & Enss, T. Excitation spectra and RF response near the polaron-to-molecule transition from the functional renormalization\ngroup. Phys. Rev. A 83, 063620 (2011).\n[25] Recati, A. & Stringari, S. Spin fluctuations, susceptibility and the dipole oscillation of a nearly ferromagnetic Fermi gas. Phys. Rev. Lett.\n106, 080402 (2011).\n[26] Bienaim ´e, T. et al. Spin-dipole oscillation and polarizability of a binary Bose-Einstein condensate near the miscible-immiscible phase\ntransition. Phys. Rev. A 94, 063652 (2016).\n[27] Sanner, C. et al. Speckle imaging of spin fluctuations in a strongly interacting Fermi gas. Phys. Rev. Lett. 106, 010402 (2011).\n[28] Sommer, A., Ku, M., Roati, G. & Zwierlein, M. Universal spin transport in a strongly interacting Fermi gas. Nature 7342 , 201 (2011).16\n[29] Enss, T. & Haussmann, R. Quantum mechanical limitations to spin diffusion in the unitary Fermi gas. Phys. Rev. Lett. 109, 195303\n(2012).\n[30] Bardon, A. B. et al. Transverse demagnetization dynamics of a unitary Fermi gas. Science 344, 722 (2014).\n[31] Burchianti, A. et al. Efficient all-optical production of large6Li quantum gases using D1gray-molasses cooling. Phys. Rev. A 90, 043408\n(2014).\n[32] Valtolina, G. et al. Josephson effect in fermionic superfluids across the BEC-BCS crossover. Science 350, 1505 (2015).\n[33] Z ¨urn, G. et al. Precise characterization of6LiFeshbach resonances using trap-sideband-resolved RF spectroscopy of weakly bound\nmolecules. Phys. Rev. Lett. 110, 135301 (2013).\n[34] Duine, R. A., Polini, M., Stoof, H. T. C. & Vignale, G. Spin Drag in an ultracold Fermi gas on the verge of ferromagnetic instability.\nPhys. Rev. Lett. 104, 220403 (2010).\n[35] Taylor, E., Zhang, S., Schneider, W., & Randeria, M. Colliding clouds of strongly interacting spin-polarized fermions. Phys. Rev. A 84,\n063622 (2011).\n[36] Nascimbene, S. et al. Fermi-Liquid behavior of the normal phase of a strongly interacting gas of cold atoms. Phys. Rev. Lett. 106, 215303\n(2011).\n[37] Tajima, H., Hanai, R. & Ohashi, Y . Strong-coupling corrections to spin susceptibility in the BCS-BEC-crossover regime of a superfluid\nFermi gas. Phys. Rev. A 93, 013610 (2016).\n[38] Goulko, O., Chevy, F. & Lobo, C. Collision of two spin-polarized fermionic clouds. Phys. Rev. A 84, 051605 (2011).\n[39] Elliott, E., Joseph, J. A. & Thomas, J. E. Anomalous minimum in the shear viscosity of a Fermi gas. Phys. Rev. Lett. 113, 020406 (2014).\n[40] Trotzky, S. et al. Observation of the Leggett-Rice effect in a unitary Fermi gas. Phys. Rev. Lett. 114, 015301 (2015).\n[41] Levinsen, J. & Parish, M. M. Strongly interacting two-dimensional Fermi gases . V ol. 3, Chap. 1, pp. 1–75 of Annual Review of Cold\nAtoms and Molecules (World Scientific, 2015).\n[42] Pilati, S., Zintchenko, I. & Troyer, M. Ferromagnetism of a repulsive atomic Fermi gas in an optical lattice: a quantum Monte Carlo\nstudy. Phys. Rev. Lett. 112, 015301 (2014).\n[43] Pilati, S. & Fratini, E. Ferromagnetism in a repulsive atomic Fermi gas with correlated disorder. Phys. Rev. A 93, 051604(R) (2016).\n[44] He, L., Liu, X.-J., Huang, X.-G. & Hu, H. Stoner ferromagnetism of a strongly interacting Fermi gas in the quasirepulsive regime. Phys.\nRev. A 93, 063629 (2016).\n[45] Zintchenko, I., Wang, L. & Troyer, M. Ferromagnetism of the Repulsive Atomic Fermi Gas: three-body recombination and domain\nformation. Preprint at: http://arxiv.org/abs/1308.1961 (2013).\n[46] Lipparini, E. & Stringari, S. Sum rules and giant resonances. Phys. Rep. 175, 103 (1989).\n[47] Nozi `eres, P. & Pines, D. The theory of quantum liquids (Perseus Ed., 1999).\n[48] Chin, C. & Julienne, P. S. Radio-frequency transitions on weakly bound ultracold molecules. Phys. Rev. A 71, 012713 (2005).\n[49] Massignan, P. & Bruun, G. M. Repulsive polarons and itinerant ferromagnetism in strongly polarized Fermi gases. Eur. Phys. J. D 65,\n83–89 (2011).\n[50] Wlazlowski, G. et al. Cooper pairing above the critical temperature in a unitary Fermi gas. Phys. Rev. Lett 110, 090401 (2013).\n[51] Koschorreck, M. et al. Attractive and repulsive Fermi polarons in two dimensions. Nature 485, 619 (2012).\n[52] Haller, E. et al. Realization of an excited, strongly correlated quantum gas phase. Science 325, 1224–1227 (2009).\n[53] Vichi, L. & Stringari, S. Collective oscillations of an interacting trapped Fermi gas. Phys. Rev. A 60, 474 (1999).\n[54] Jeon, J. W. & Mullin, W. J. Theory of spin diffusion of dilute, polarized fermions for all temperatures. J. Low Temp. Phys. 67, 421 (1987).\n[55] Bruun, G. M. Spin diffusion in Fermi gases. New J. Phys. 13, 035005 (2011).\n[56] Enss, T. Transverse spin diffusion in strongly interacting Fermi gases. Phys. Rev. A 88, 033630 (2013).\n[57] Bruun, G. M., Recati, A., Pethick, C. J., Smith, H. & Stringari, S. Collisional properties of a polarized Fermi gas with resonant\ninteractions. Phys. Rev. Lett 100, 240406 (2011).\n[58] Bishop, R. F., Ghassib, H. B. & Strayer, M. R. Composite pairs and effective two-body scattering in a many-body medium. Phys. Rev. A\n13, 1570 (1976).\n[59] Bruun, G. M. & Smith, H., Viscosity and thermal relaxation for a resonantly interacting Fermi gas. Phys. Rev. A 72, 043605 (2005).17\n[60] Chiacchiera, S., Lepers, T., Davesne, D. & Urban, M. Collective modes of trapped Fermi gases with in-medium interaction. Phys. Rev.\nA79, 033613 (2009).\n[61] Enss, T. Quantum critical transport in the unitary Fermi gas. Phys. Rev. A 86, 013616 (2012).\n[62] Bluhm, M. & Sch ¨afer, T. Medium effects and the shear viscosity of the dilute Fermi gas away from the conformal limit. Phys. Rev. A 90,\n063615 (2014).\n[63] Gensemer, S.D. & Jin, D.S. Transition from collisionless to hydrodynamic behavior in an ultracold Fermi gas. Phys. Rev. Lett. 87, 173201\n(2001)." }, { "title": "1502.04934v1.Direct_Observation_of_Ferromagnetic_State_in_Gold_Nanorods_Probed_using_Electron_Spin_Resonance_Spectroscopy.pdf", "content": "arXiv:1502.04934v1 [cond-mat.mes-hall] 17 Feb 2015Direct Observation of Ferromagnetic State in Gold Nanorods\nProbed using Electron Spin Resonance Spectroscopy\nYuji Inagaki∗and Tatsuya Kawae\nDepartment of Applied Quantum Physics, Faculty of Engineer ing,\nKyushu University, Fukuoka 819-0395, Japan\nNatsuko Sakai and Yuta Makihara\nDepartment of Material Physics and Chemistry,\nFaculty of Engineering, Kyushu University, Fukuoka 819-03 95, Japan\nHiroaki Yonemura and Sunao Yamada\nDepartment of Applied Chemistry, Faculty of Engineering,\nKyushu University, Fukuoka 819-0395, Japan\nAbstract\nX-band electron spin resonance (ESR) spectroscopy has been performed for gold nanorods\n(AuNRs) of four different sizes covered with a diamagnetic sta bilizing component, cetyltrimethy-\nlammmonium bromide. The ESR spectra show ferromagnetic fea tures such as hysteresis and reso-\nnance field shift, depending on the size of the AuNRs. In addit ion, the ferromagnetic transition is\nindicated by an abrupt change in the spectra of the two smalle st AuNRs studied. A large g-value\nin the paramagnetic region suggests that the ferromagnetis m in the AuNRs originates from strong\nspin-orbit interaction.\n1I. INTRODUCTION\nIn the past decade, rod-shaped gold nanoparticles, called Au nano rods (AuNRs), have\nattracted much attention because of their potential applications in sensing, imaging, and in\nvivo photothermal cancer therapy1–4. These applications are based on the inherent tunable\noptical properties of the AuNRs by changing the aspect ratio ardefined by the ratio of the\nlength to the diameter5. In contrast, there have been few reports thus far concerning the\nmagnetic properties of AuNRs6, although there have been many studies on the magnetism\nof spherical Au nanoparticles (AuNPs)7–12.\nRecently, Yonemura et al.examined the effects of magnetic processing and observed\nthe magnetic orientation and the side-by-side aggregation of AuNR s/poly(styrenesulfonate)\ncomposites, AuNRs with three different aspect ratios ( ar= 8.3, 5.1, and 2.5), and Au\nnanowires under a magnetic field13–15, which implies that the magnetic response of AuNRs\nis positive, although bulk Au has been believed to be diamagnetic for a lo ng time. However,\nthe detailed properties and origin of magnetism in AuNRs is not yet well understood.\nGenerally, magnetic properties have been investigated through ma gnetization measure-\nments using highly sensitive magnetometers, such as a commercial S QUID magnetometer.\nIn nano-sized particles coated with a stabilizing polymer, however, it is difficult to extract\nthe net magnetization precisely owing to the superposition of a large amount of diamagnetic\ncomponents16. We focus on the electron spin resonance (ESR) technique to elimina te the\nsignal from the stabilizing polymer because ESR is insensitive to diamag netism. Moreover,\nthe skin depth of Au at the X-band microwave region ( ∼9 GHz) is about 0.8 µm, which is\nmuch larger than the size of AuNRs in the present study. These fea tures indicate that ESR\nis a powerful tool for investigating the magnetic properties of AuN Rs.\nIn the present study, X-band ESR measurements were performe d in detail for AuNRs\nof four different sizes, where arwas varied from 2.5 to 8.3. Ferromagnetic features were\nobserved in all AuNRs. In addition, ferromagnetic transitions were clearly detected in the\ntwo smallest AuNRs studied. From the ESR parameters deduced fro m the conduction\nelectron spin resonance (CESR) at high temperatures, strong sp in-orbit interaction and\nresultant large effective mass of conduction electrons are pointed out for the origin of the\nmagnetism in the AuNRs on the basis of Elliott-Yafet theory.\n2TABLE I. Dimensions, aspect ratio, estimated content of Au, CTAB and other metals for each\nAuNR\nsize (nm) arAu(%) CTAB(%) Ag(%)\ns-AuNR 4.6 φ×11.6 2.5 3.6 Br:22.6, C:59.3, H:10.8, N:3.7 -\ns′-AuNR 5.0 φ×20.0 4.0 41.3 Br:16.7, C:31.4, H:5.8, N:1.9 2.9\nm-AuNR 7.2 φ×36.6 5.1 3.1 Br:22.9, C:59.4, H;11.0, N:3.6 -\nl-AuNR 7.7 φ×63.8 8.3 1.9 Br:22.7, C:60.6, H:11.2, N:3.7 -\nII. EXPERIMENTAL\nAuNRs studied here are named as s-, s′-, m-, and l-AuNR from small to large values\nofar. All AuNRs are prepared by the soft template method using cetyltr imethylamm-\nmonium bromide (CTAB)17. The s- and s′-AuNR are prepared using a reducing agent,\ntriethylamine, with or without acetone. The m-AuNR is prepared usin g a combination of\nchemical reduction and photoreduction. The l-AuNR is prepared us ing two kinds of re-\nducing agents, sodium borohydride and triethylamine. Typical TEM im ages of AuNRs are\ngiven in ref.13. X-ray fluorescence and CHN elemental analyses confirm that AuN Rs contain\nno contamination of other magnetic metals, as listed in Table I, which e nsures that the\nmagnetic response in the present study originates from the Au ato ms in AuNRs. ESR mea-\nsurements are performed using an X-band microwave system (JEO L ES-SCEX) equipped\nwith a continuous-He-flow-type cryostat (Oxford ESR910) oper ating down to T∼5 K.\nIII. RESULTS AND DISCUSSION\nFigures 1(a) and 1(b) show the temperature dependence of ESR s pectra observed for\ns-AuNR and s′-AuNR, respectively. Two spectra plotted with gray lines in Fig. 1(a) corre-\nspond to the blank signals recorded at 7 K and 295 K, which originate f rom the background\nfrom instruments including the sample holder (quartz tube) and the CTAB. This indicates\nthat the absorptions at around 320 mT are not intrinsic signals from the AuNR. At a glance,\nit seems that the spectra from the two AuNRs show quite similar temp erature dependences.\nHowever, on closely observing both spectra, several differences can be discerned. At room\ntemperature, the ESR spectrum of s-AuNR consists of two compo nents: a sharp absorption\n30 100200300400500600700 \nH (mT) 286 K \n200 K \n100 K \n6 K \n(d) 0 100 200 300 400 500 600 \nH (mT) (b) 297 K \n251 K \n202 K \n150 K \n120 K \n100 K \n80 K \n62 K \n49 K \n41 K \n30 K \n21 K \n12.5 K \n5.5 K \n0 100 200 300 400 500 Derivative Intensity (arb. units) \nH (mT) 295 K \n200 K \n140 K \n103 K \n75 K \n60 K \n51 K \n40 K \n32 K \n16.5 K \n9.7 K \n6.0 K \nblank 7 K blank 295 K \n(a) \n0 100 200 300 Derivative intensity (arb.units) \nH (mT) 300 K \n250 K \n200 K \n100 K \n88 K \n75 K \n62 K \n45 K \n32 K \n15 K \n10 K \n6.8 K 150 K \n(c) \nFIG. 1. (Color online) Temperature dependence of ESR spectr a in (a) s-AuNR, (b) s′-AuNR (c)\nm-AuNR and (d) l-AuNR. The absorption around 320 mT is due to i nstrumental background, as\nobserved from the blank spectra recorded at 295 K and 7 K shown by the gray lines in (a).\ncentered at s 1= 275 mT and a broad one around s 2= 230 mT, which is strongly suppressed\nwhen the temperature is decreased. In contrast, the spectra o f s′-AuNR have no broad\ncomponent and show only a sharp absorption centered at s′\n1= 289 mT. Thus, the sharp ab-\nsorptions at s 1and s′\n1are regarded as intrinsic properties of AuNRs, and their temperat ure\nvariations are examined below.\nThe temperature dependences of g-values estimated from resonance fields and line widths\n∆Hfor s1and s′\n1are summarized in Fig. 2. The resonance field and line width are nearly\nindependent of temperature for both samples above ∼80 K, while the spectra show a drastic\nshift of the resonance field and broadening of the width below 55 K fo r s-AuNR and 75\n41.5 2.0 2.5 3.0 \n020 40 60 80 100 \n0 50 100 150 200 250 300 g-value ΔH (mT) \nT (K) TcsTcs' \nFIG. 2. Temperature dependence of g-value (circle) and line width ∆ H(square). Open and solid\nsymbols represent the results for s-AuNR and s′-AuNR, respectively.\nK for s′-AuNR. The g-value of s-AuNR (s′-AuNR) increases with decreasing temperature\nand reaches 3.19 (3.07) at the lowest temperature T∼6 K. For s-AuNR (s′-AuNR), the\nline width shows a maximum at approximately 40 K (60 K), which is followed by a gradual\nnarrowing with decreasing temperature. This series ofbehaviors is atypical featureobserved\nin the magnetic ordering process. Hence, we conclude that the res onance field shift and\nbroadening of the width in s- and s′-AuNRs are caused by magnetic ordering with transition\ntemperatures of Ts\nc∼55 K and Ts′\nc∼75 K, respectively.\nThe results for m- and l-AuNRs are depicted in Figs. 1(c) and 1(d), r espectively. In\nm-AuNR, the ESR spectrum cannot be observed in the high-temper ature region. When the\ntemperature is decreased below T∼100 K, a narrow absorption with ∆ H= 8 mT appears\nat around 150 mT. The intensity of the spectrum grows with decrea sing temperature, while\nboth the resonance center and line width are almost independent of temperature down to T\n= 6.8 K. This temperature dependence will be discussed later.\nIn contrast, l-AuNR shows broad ESR spectra in the entire temper ature range, as shown\nin Fig. 1(d). The resonance field at ∼70mT in l-AuNR is the lowest among all AuNRs, while\nthe line width of ∼130 mT is the broadest. It is significant that a hysteresis emerges in all\nthe spectra of l-AuNR below ∼60 mT in the magnetic field sweep, as indicated by arrows. A\nhysteresis between magnetizing and demagnetizing processes is a c haracteristic feature of a\n50 50 100 150 200 250 300 350 400 Derivative intensity (arb. units) \nH (mT) g = 2 \nFIG. 3. (Color online) ESR spectra of all AuNRs at T ∼6K. Red, blue, green and orange colors\ncorrespond to s-, s′-, m- and l-AuNR, respectively. The vertical dotted line ind icates the resonance\nfield forg=2.\nferromagnet with domain structures and/or magnetic anisotropy18,19. Note that the domain\nstructure and anisotropy give rise to the line broadening in ESR spec tra. In other words,\nthese features observed in l-AuNR are well explained by assuming a la rge scale ferromagnet.\nNext, we discuss the systematic resonance shift in all AuNRs. The s pectra at the lowest\ntemperature T∼6 K are plotted together in Fig. 3 to make a comparison between all\nAuNRs. In a ferromagnet with a cylindrical shape like the present Au NRs, ferromagnetic\nresonance (FMR) occurs at a resonance field that depends on the demagnetizing form factor\nowing to the aspect ratio ar, as well as on the magnitude of moment Msand anisotropy K.\nIn fact, the magnitude of shift increases with ar in the present systems; the shift is estimated\nto be approximately 248, 176, 114, and 118 mT for l-, m-, s′-, and s-AuNR, respectively,\nwhich corresponds to the order of arexcept for s- and s′-AuNR. The reversal between them\nmay be caused by the resonance shift dominated by an effective anis otropy field represented\nbyK/Ms. Thus, a small anisotropy Kcan give rise to large effective field if Msis small. As\nfor the anisotropy, the following qualitative discussion can be made. In randomly oriented\nferromagnetic species, a powder pattern is expected in the FMR sp ectrum. Although a\nclear powder pattern was not recorded for all AuNRs, finite ESR int ensity can be seen like\na tail in a higher field range than the main absorption peak, as repres ented in l-AuNR. It\nis difficult to refer about other AuNRs because of the superposition of background signal\nat around 320 mT. However, a small intensity like shoulder is visible at a bout 270 mT in\n6s′-AuNR, as indicated by the vertical arrow in Fig. 3. Such FMR spectr a correspond to the\ncase with a negative anisotropy in the cubic symmetry20. Detailed frequency dependence of\nESR is required to obtain further information about the FMR parame tersMsandK.\nThe observed systematic shift, which is generally not expected in ot her antiferromagnets,\nparamagnets, and ferromagnets with a spherical shape, provide s further evidence to the\nexistence of ferromagnetic states at low temperatures in all the p resent AuNRs. On the\nbasis of the hysteresis and systematic shift of the resonance field , it is reasonable to consider\nthat all AuNRs are in the ferromagnetic state at low temperatures . To our knowledge, this\nis the first time that a ferromagnetic state has been found in AuNRs .\nIn the final part of this Letter, we examine the ESR spectra in the p aramagnetic region.\nThe spectra of s- and s′-AuNRs above Tccan be understood as CESR ones. In the CESR\nregion,g-values are estimated to be 2.34 ±0.09 and 2.26 ±0.06 for s- and s′-AuNRs, respec-\ntively. These values are considerably larger than the value of 2.11 re ported for bulk Au\nbut comparable to that of 2.26 ±0.02 for small particles of Au with a mean diameter of 3\nnm21,22. For CESR, the difference ∆ gbetween the g-value of real metals and that of ideal\nfree electrons (2.0023) gives an approximate value of the spin-orb it interaction through the\nequation ∆ g∼λ/∆E. Here,λis the spin-orbit coupling constant and ∆ Eis the difference in\nenergy between the 6s band and the nearest 5d band23. Thus, the large ∆ gobserved in the\npresent measurements indicates the large contribution of the orb ital moment in the 5d band\nto conduction electrons in the 6s band, which governs the magnetic properties of AuNRs\nand leads to the ferromagnetic state at low temperatures. A stro ng spin-orbit coupling was\nalso confirmed in recent studies of X-ray magnetic circular dichroism in bulk Au as well as\nAuNPs24,25.\nThe strong spin-orbit interaction also causes a significant broaden ing of line width. The\nline width of CESR in metals is closely related to the spin-lattice relaxatio n time, i.e., the\nspin flip rate by phonons. Accordingly, ∆ Hvaries linearly with temperature. Therefore,\nat high temperatures, CESR is hardly detected in not only bulk metals but also m-AuNR.\nThis is a reason why we could not observe the ESR in m-AuNR above 100 K. In contrast,\nnarrow ESR absorptions are obtained for both s- and s′-AuNRs, which show rather gentle\ntemperature dependences without a marked increase in the width. These features may be\nexplained by considering the system size of AuNRs. In small systems , there exists a lower\nlimit of phonon mode frequency given by ν=vs/2L, whereLis the largest dimension of\n710 510 610 710 810 910 10 10 11 10 12 10 13 \n10 -7 10 -6 10 -5 10 -4 10 -3 10 -2 10 -1 10 0ΔH/ρ (mT/ Ω/m) \nΔg2Au \nPd \nNa KRb Cs \nCu Ag \nAl \ns'-AuNR \ns-AuNR \nFIG. 4. Beuneu-Monod plot, which connects ∆ g2and ∆H/ρ, for pure metals given in ref.27and\nfor two systems in the present study. The value ∆ H/ρof most metals differs from the fit, which\nis indicated by the solid and dashed lines, by less than one or der of magnitude.\nthe system and vsis the speed of sound in the crystal; this results in an increase in the\nspin-lattice relaxation time of conduction electrons26. As a result, narrow absorptions are\nobserved in s- and s′-AuNRs. This scenario is qualitative and does not account for the\nnearly equal line widths of s-AuNR and s′-AuNR; other factors should be considered for the\nquantitative explanation.\nNevertheless, owing to the narrow line widths, we could successfully observe CESR of\nAuNR, which allows identification of the characteristic feature of th e present CESR results\nthrough sorting with other metals in the so-called BeuneuMonod plot shown in Fig. 427,28.\nThe BeuneuMonod plot is an empirical plot that connects CESR param eters ∆gand ∆H\nfor pure metals via ∆ H/ρ=α∆g2, whereρis the resistivity and αis a metal-dependent\nconstant. Most metals follow a straight-line fit in the log-log plot of ∆ H/ρvs. ∆g2by less\nthan one order of magnitude, as indicated by the solid and dashed line s in Fig. 4. The values\nused in this plot are taken at a temperature of approximately TD/7, where TDis the Debye\ntemperature. In the case of bulk Au, TD/7 corresponds to 20 K. The CESR parameters\nfor the two systems in the present study (s- and s′-AuNR) are not available for such a low\ntemperature, which is less than Tc. Thus, we plot the estimate taken in the temperature\n8range between Tcand room temperature, while the resistivity is fixed at the bulk value a t 20\nK. Error bars for both axes originate from the temperature varia tion of ∆ gand ∆H. In the\nplot, the two data points corresponding to s- and s′-AuNR deviate significantly fromthose of\nother metals, but they are located near the point corresponding t o Pd. It is well known that\nPd is close to satisfying the Stoner criterion even in a bulk form. Inde ed, ferromagnetism is\nrealized by the downsizing of Pd29–31. Accordingly, it is confirmed that the two AuNRs are\nalso close to the ferromagnetic state.\nTo reproduce the results for s-AuNR, s′-AuNR, and Pd, a smaller αis needed. This\nsuggests that the low value of αis closely related to the realization of ferromagnetism.\nAccording to the Elliott-Yafet theory, the coefficient αis given by ne2/γm∗, wheren,e, and\nm∗are the density, charge, and effective mass of the conduction elec tron, respectively, and γ\nis the magnetomechanical ratio. Therefore, the origin of the devia tion caused by the narrow\n∆Hand the large ∆ gis suggested to be an enhancement of the effective mass. This is a\nreasonable conclusion because a large effective mass is realized in a ty pical ferromagnet of\nNi, as revealed recently by high-resolution angle-resolved photoem ission studies32.\nIn summary, we performed ESR measurements for AuNRs of four s izes with different\naspect ratios. The ESR spectra at low temperatures are explained in the context of ferro-\nmagnetic resonance for all AuNRs. Detailed frequency dependenc e of FMR measurements\nwill enable further quantitative discussion on the size- and shape-d ependent magnetism of\nAuNRs. In the two smallest AuNRs, we detected ferromagnetic tra nsitions in the ESR spec-\ntra atT∼60 K, which offers an opportunity to explore the critical behavior of the phase\ntransition in nano-rod systems. The CESR above Tcsuggests that the strong spin-orbit\ninteraction is responsible for ferromagnetism in the AuNR systems.\nACKNOWLEDGMENTS\nThe authors are grateful to T. Asano and T. Sakurai for the help of ESR experiments\nand useful discussion. This work was partially supported by a Grant -in-Aid for Scientific\nResearch, No.23540392, No.25220605 and No.25287076. The auth ors thank to Dr. Daigou\nMizoguchi (Dai Nippon Toryo Co. Ltd.) for providing them with four d ifferent types of\nAuNRs. The authors are also grateful to the Center of Advanced Instrumental Analysis,\nKyushu University for X-ray fluorescence, CHN elemental analyse s and the use of a TEM\n9apparatus.\n∗inagaki.yuji.318@m.kyushu-u.ac.jp\n1C. S¨ onnichsen, T. Franzl, T. Wilk, G. von Plessen, J. Feldma nn, O. Wilson, and P. Mulvaney,\nPhys. Rev. Lett. 88, 077402 (2002).\n2K. K. Caswell, J. N. Wilson, U. H. F. Bunz, and C. J. Murphy, J. A n. Chem. Soc. 125, 13914\n(2003).\n3K. G. Thomas, S. Barazzouk, B. I. Ipe, S. S. Joseph, and P. V. Ka mat, J. Phys. Chem. B 108,\n13066 (2004).\n4X. Huang, I. El-Sayed, and M. A. El-Sayed, Cancer Nanotechno logy: Method and Protcols\n624, 343 (2010).\n5A. Brioude, X. C. Jiang, and M. P. Pileni, J. Phys. Chem. 109, 13138 (2005).\n6P. G. van Rhee, P. Zijlstra, T. G. A. Verhagen, J. Aarts, M. I. K atsnelson, J. C. Maan, M. Orrit,\nand P. C. M. Christianen, Phys. Rev. Lett. 111, 127202 (2013).\n7P. Crespo, R. Litr´ an, T. C. Rojas, M. Multigner, J. M. de la Fu ente, J. C. S´ anchez-L´ opez, M. A.\nGarc´ ıa, A. Hernando, S. Penad´ es, and A. Fern´ andez, Phys. Rev. Lett. 93, 087204 (2004).\n8H. Hori, Y. Yamamoto, T. Iwamoto, T. Miura, T. Teranishi, and M. Miyake, Phys. Rev. B 69,\n174411 (2004).\n9J. S. Garitaonandia, M. Insausti, E. Goikalea, M. Suzuki, J. D. Cashion, N. Kawamura, H. Oh-\nsawa, I. G. de Muro, K. Suzuki, F. Plazoala, and T. Rojo, Nano L etters8, 661 (2008).\n10C.-Y. Li, C.-M. Wu, S. K. Karna, C.-W. Wang, D. Hsu, C.-J. Wang , and W.-H. Li, Phys. Rev.\nB83, 174446 (2011).\n11R. Gr´ eget and M. Others, ChemPhysChem 13, 3092 (2012).\n12Y.Negishi, H.Tsunoyama, M.S.N.Kawamura, M. M. Matsushita , K.Maruyama, T.Sugawara,\nT. Yokoyama, and T. Tsukuda, J. Am. Chem. Soc. 128, 12034 (2006).\n13H. Yonemura, J. Suyama, T. Arakawa, and S. Yamada, Thin Solid Films518, 668 (2009).\n14H. Yonemura, N. Sakai, J. Suyama, and S. Yamada, Mol. Cryst. L iq. Cryst. 599, 63 (2014).\n15H. Yonemura, Y. Makihara, N. Sakai, M. Iwasaka, and S. Yamada , J. PhotoChem. Photobil.\nA220, 179 (2011).\n16M. A. Garcia, E. F. Pinel, J. dela Venta, A. Quesada, V. Bouzas , J. F. Fern´ andez, J. J. Romero,\n10M. S. M. Gonz´ alez, and J. L. Costa-Kr¨ amer, J. Appl. Phys. 105, 013925 (2009).\n17Y. Niidome, K. Nishioka, H. Kawasaki, and S. Yamada, Chem. Co mmun.18, 2376 (2003).\n18D. L. Griscom, C. L. Marquardt, and E. J. Friebele, Earth Plan et. Sic. Lett. 24, 78 (1974).\n19V. S. Bhat, J. Sklenar, B. Farmer, J. Woods, J. T. Hastings, S. J. Lee, J. B. Ketterson, and\nL. E. Long, Phys. Rev. Lett. 111, 077201 (2013).\n20G. V. Skrotskii and L. V. Kurbatov, in Ferromagnetic Resonance , edited by S. V. Vonsovskii\n(Pergamon Press, 1966) Chap. II, pp. 30–44.\n21P. Monod and A. Janossy, J. Low Temp. Phys. 26, 311 (1977).\n22R. Dupree, C. T. Forwood, and M. J. A. Smith, Phys. Stat. Sol. 24, 525 (1967).\n23Y. Yafet, Solid State Phys. 14, 1 (1963).\n24M. Suzuki, N. Kawamura, H. Miyagawa, J. S. Garitaonandia, Y. Yamamoto, and H. Hori,\nPhys. Rev. Lett. 108, 047201 (2012).\n25Y.Yamamoto, T.Miura, M. Suzuki, K.N, H. Miyagawa, T.Nakamu ra, K.Kobayashi, T.Teran-\nishi, and H. Hori, Phys. Rev. Lett. 93, 116801 (2004).\n26A. Honig, in Paramagnetic Resonance , Vol. II, edited by W. Low (Academic Press, New York,\nLondon, 1963) pp. 439–442.\n27F. Beuneu and P. Monod, Phys. Rev. B 18, 2422 (1978).\n28P. Monod and F. Beuneu, Phys. Rev. B 19, 911 (1979).\n29T. Taniyama, E. Ohta, and T. Sata, Europhys. Lett. 38, 195 (1997).\n30T. Shinohara, T. Sato, and T. Taniyama, Phys. Rev. Lett. 91, 197201 (2003).\n31K. Ienaga, N. Nakashima, Y. Inagaki, H. Tsujii, T. Kimura, an d T. Kawae, Appl. Phys. Lett.\n101, 123114 (2012).\n32M. Higashiguchi, K. Shimada, K. Nishiura, X. Cui, H. Namatam e, and M. Taniguchi, Phys.\nRev. B72, 214438 (2005).\n11" }, { "title": "1106.3519v1.Current_effect_on_magnetization_oscillations_in_a_ferromagnet___antiferromagnet_junction.pdf", "content": "arXiv:1106.3519v1 [cond-mat.mtrl-sci] 17 Jun 2011Current effect on magnetization oscillations\nin a ferromagnet–antiferromagnet junction\nE. M. Epshtein∗, Yu. V. Gulyaev, P. E. Zilberman,\nV. A. Kotelnikov Institute of Radio Engineering and Electronics\nof the Russian Academy of Sciences, Fryazino, 141190, Russia\nAbstract\nSpin-polarized current effect is studied on the static and dy namic mag-\nnetization of the antiferromagnet in a ferromagnet–antife rromagnet junc-\ntion. The macrospin approximation is generalized to antife rromagnets.\nCanted antiferromagnetic configuration and resulting magn etic moment\nare induced by an external magnetic field. The resonance freq uency and\ndamping are calculated, as well as the threshold current den sity corre-\nsponding to instability appearance. A possibility is shown of generating\nlow-damping magnetization oscillations in terahertz rang e. The fluctua-\ntion effect is discussed on the canted antiferromagnetic con figuration.\n1 Introduction\nThe discovery of the spin transfer torque effect in ferromagn etic junctions\nunder spin-polarized current [1, 2] has stimulated a number of works in\nwhich such effects were observed as switching the junction ma gnetic con-\nfiguration [3], spin wave generation [4], current-driven mo tion of magnetic\ndomainwalls [5], modification offerromagnetic resonance [ 6], etc. Itiswell\nknown that spin torque transfer from spin-polarized electr ons to lattice\nleads to appearance of a negative damping. At some current de nsity, this\nnegative damping overcomes the positive (Gilbert) damping with occur-\nring instability of the original magnetic configuration. Th e corresponding\ncurrent density is high enough, of the order of 107A/cm2. This, naturally,\nstimulates attempts to lower this threshold. Various ways w ere proposed,\nsuch as using magnetic semiconductors [7], in which the thre shold cur-\nrent density can be lower down to 105–106A/cm2because of their low\nsaturation magnetization. However, using of such material s requires, as\na rule, low temperatures because of low Curie temperature. B esides, the\nferromagnetic resonance frequency is rather low in this cas e.\n∗E-mail: epshtein36@mail.ru\n1FM AFMNM\nMFH−+n\nxyz\n0 LAFMj/e\nFigure 1: Scheme of the ferromagnet (FM)–antiferromagnet (AF M) junction;\nNM being a nonmagnetic layer. The main vector directions are shown.\nIn connection with these difficulties, the other approaches w ere pro-\nposed, based on high spin injection [8] or joint action of ext ernal magnetic\nfield and spin-polarized current [9, 10]. It seems promising , also, using\nmagnetic junction of ferromagnet–antiferromagnet type, i n which the fer-\nromagnet (FM) acts as an injector of spin-polarized electro ns. The anti-\nferromagnetic (AFM) layer, in which the magnetic sublattic es are canted\nby external magnetic field, may have very low magnetization t hat pro-\nmotes low threshold [11]. The AFM resonance frequency may be both\nlow and high reaching 1012s−1, i.e. terahertz (THz) range. However,\ninvestigation and application of THz resonances is prevent ed because of\ntheir large damping. Such a damping in ferromagnetic juncti ons can be\nsuppressed, as mentioned above, by means of spin-polarized current. The\nquestion arises about possibility of such a suppression in F M–AFM junc-\ntions. Note, that this problem has been paid attention of a nu mber of\nauthors [12]–[20].\n2 The equations of motion\nLet us consider a FM–AFM junction (Fig. 1) with current flowin g per-\npendicular to layers, along xaxis. An external magnetic field is parallel\nto the FM magnetization and lies in the layer plane yz. The simplest AF\nmodel is used with two equivalent sublattices.\nThe AFM energy (per unit area), with uniform and nonuniform e x-\nchange, anisotropy, external magnetic field, demagnetizat ion and the sd\nexchange interaction of the conduction electrons with the m agnetic lattice\ntaking into account, takes the form [21]\n2W=/integraldisplayLAFM\n0dx/braceleftBigg\nΛ(M1·M2)+1\n2α/braceleftBigg/parenleftbigg∂M1\n∂x/parenrightbigg2\n+/parenleftbigg∂M2\n∂x/parenrightbigg2/bracerightBigg\n+α′/parenleftbigg∂M1\n∂x·∂M2\n∂x/parenrightbigg\n−1\n2β/braceleftbig\n(M1·n)2+(M2·n)2/bracerightbig\n−β′(M1·n)(M2·n)\n−((M1+M2)·H)−αsd((M1+M2)·m)+2π(M1+M2)2\nx/bracerightBigg\n,(1)\nwhereM1,M2are the sublattice magnetization vectors, Λ is the uni-\nform exchange constant, α, α′are the intrasublattice and intersublattice\nnonuniform exchange constants, respectively, β, β′are the corresponding\nanisotropy constants, nis the unit vector along the anisotropy axis, His\nthe external magnetic field, mis the conduction electron magnetization,\nαsdis the dimensionless sdexchange interaction constant; the last term\ndescribes demagnetization effect. The integral is taken ove r the AFM\nlayer thickness LAFM. We are interested in the spin-polarized current\neffect on the AFM layer, so we consider a case of perfect FM inje ctor with\npinned lattice magnetization and without disturbance of th e electron spin\nequilibrium, that allows to not include the FM layer energy i n Eq. (1).\nTwo mechanisms are known of the spin-polarized current effec t on\nthe magnetic lattice, namely, spin transfer torque (STT) [1 , 2] and an\nalternative mechanism [22, 23] due to the spin injection and appearance of\nnonequilibrium population of the spin subbands in the colle ctor layer (this\nis AFM layer, in our case). In the case of antiparallel relati ve orientation\nof the injector and collector magnetization vectors, such a state becomes\nenergetically unfavorable, so that the antiparallel config uration switches\nto parallel one (such a process in FM junction is considered i n detail in\nreview [24]). The latter mechanism is described with the sdexchange\nterm in Eq. (1). As to the former mechanism, it is of dissipati ve character\n(it leads to negative damping), so that it is taken into accou nt by the\nboundary conditions (see below), not the Hamiltonian.\nThe equations of the sublattice motion with damping taking i nto ac-\ncount take the form\n∂Mi\n∂t−κ\nM0/bracketleftbigg\nMi×∂Mi\n∂t/bracketrightbigg\n+γ/bracketleftBig\nMi×H(i)\neff/bracketrightBig\n= 0 (i= 1,2),(2)\nwhereM0is the sublattice magnetization, κis the damping constant,\nH(i)\neff=−δW\nδMi(i= 1,2) (3)\nare the effective fields acting on the corresponding sublatti ces.\nFrom Eqs. (1)–(3) the equations are obtained for the total ma gnetiza-\ntionM=M1+M2and antiferromagnetism vector L=M1−M2:\n3∂M\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nM×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nL×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[M×H]+γ[M×Hd]+γ[M×Hsd]\n+1\n2γ(β+β′)(M·n)[M×n]+1\n2γ(β−β′)(L·n)[L×n]\n+1\n2γ(α+α′)/bracketleftbigg\nM×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nL×∂2L\n∂x2/bracketrightbigg\n= 0,(4)\n∂L\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nL×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nM×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[L×H]+γ[L×Hd]+γ[L×Hsd]−γΛ[L×M]\n+1\n2γ(β+β′)(M·n)[L×n]+1\n2γ(β−β′)(L·n)[M×n]\n+1\n2γ(α+α′)/bracketleftbigg\nL×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nM×∂2L\n∂x2/bracketrightbigg\n= 0,(5)\nwhereHd=−4π{M1x+M2x,0,0}is the demagnetization field,\nHsd(x) =δ\nδM(x)/integraldisplayLAFM\n0dx′/parenleftbig\nM(x′)·m(x′)/parenrightbig\n(6)\nis the effective field due to sdexchange interaction. This field determines\nthe spin injection contribution to the interaction of the co nduction elec-\ntrons with the antiferromagnet lattice.\nTo findHsd(x) field, the conduction electron magnetization m(x) is\nto be calculated. The details of such calculations are prese nted in our\npreceding papers [25, 9]. Here we adduce the result for the ca se, where\nthe antiferromagnet layer thickness LAFMis small compared to the spin\ndiffusion length lwith the current flow direction corresponding to the\nelectron flux from FM to AFM:\nm= (m+∆m)ˆM,∆m=µBτQj\neLAFM/parenleftBig\nˆM(0)·ˆMF/parenrightBig\n, (7)\nwheremis the equilibrium (in absence of current) electron magneti zation,\n∆mis the nonequilibrium increment due to current, ˆM=M/|M|is the\nunit vector along the AFM magnetization, ˆMFis the similar vector for\nFM,µBis the Bohr magneton, eis the electron charge, τis the electron\nspin relaxation time, jis the current density.\nIt should have in mind in varying the integral (6), that the el ectron\nmagnetization mdepends on the vector Morientation relative to the FM\nmagnetization vector MF. From Eqs. (6)and (7) we have [9]\nHsd=αsdmˆM+αsdµBτQj\neLAFMˆM+αsdµBτQj\neˆMFδ(x−0).(8)\nBy substitution (8) into (4) and (5), we obtain\n4∂M\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nM×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nL×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[M×H]+γ[M×Hd]+γαsdµBτQj\ne/bracketleftBig\nM׈MF/bracketrightBig\nδ(x−0)\n+1\n2γ(β+β′)(M·n)[M×n]+1\n2γ(β−β′)(L·n)[L×n]\n+1\n2γ(α+α′)/bracketleftbigg\nM×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nL×∂2L\n∂x2/bracketrightbigg\n= 0,(9)\n∂L\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nL×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nM×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[L×H]+γ[L×Hd]+γαsdµBτQj\ne/bracketleftBig\nL׈MF/bracketrightBig\nδ(x−0)\n−γ/parenleftbigg\nΛ−αsdm\nM−αsdµBτQj\neLAFMM/parenrightbigg\n[L×M]\n+1\n2γ(β+β′)(M·n)[L×n]+1\n2γ(β−β′)(L·n)[M×n]\n+1\n2γ(α+α′)/bracketleftbigg\nL×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nM×∂2L\n∂x2/bracketrightbigg\n= 0.(10)\n3 The boundary conditions\nThe equations of motion (9) and (10) contain derivative over the space\ncoordinate x. Therefore, boundary conditions at the AFM layer surfaces\nx= 0 and x=LAFMare need to find solutions. The way of derivation\nwas described in Ref. [9] in detail. The conditions depend on the electron\nspin polarization and are determined by the continuity requ irement of the\nspin currents at the interfaces.\nThe terms with the space derivative in Eq. (9) may be written i n the\nform of a divergency:\n1\n2γ(α+α′)/bracketleftbigg\nM×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nL×∂2L\n∂x2/bracketrightbigg\n=∂\n∂x/braceleftbigg1\n2γ(α+α′)/bracketleftbigg\nM×∂M\n∂x/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nL×∂L\n∂x/bracketrightbigg/bracerightbigg\n≡∂JM\n∂x. (11)\nTheJMvector is the lattice magnetization flux density.\nLet us integrate Eq. (9) over xwithin narrow interval 0 < x < ε with\nsubsequent passing to ε→+0 limit. Then only the mentioned terms with\nthe space derivative and the singular term with delta functi on will con-\ntribute to the integral. As a result, we obtain an effective ma gnetization\nflux density with sdexchange contribution at the AFM boundary x= +0\ntaking into account:\nJeff(+0) =JM(+0)+γαsdµBτQj\ne/bracketleftBig\nM(+0)׈MF/bracketrightBig\n.(12)\n5The magnetization flux density coming from the FM injector is\nJ(−0) =µBQ\nejˆMF. (13)\nThe component J/bardbl=/parenleftBig\nJ(−0)·ˆM(+0)/parenrightBig\nˆM(+0) remains with the elec-\ntrons, while the rest,\nJ⊥=J(−0)−J/bardbl=µBQ\nej/braceleftBig\nˆMF−ˆM(+0)/parenleftBig\nˆMF·ˆM(+0)/parenrightBig/bracerightBig\n=−µBQ\neM2j/bracketleftBig\nM(+0)×/bracketleftBig\nM(+0)׈MF/bracketrightBig/bracketrightBig\n, (14)\nis transferred to the AFM lattice owing to conservation of th e magnetiza-\ntion fluxes [1, 2].\nBy equating the magnetization fluxes (12) and (14), we obtain\nJM=−µBQ\neM2j/bracketleftBig\nM×/bracketleftBig\nM׈MF/bracketrightBig/bracketrightBig\n−γαsdµBτQ\nej/bracketleftBig\nM׈MF/bracketrightBig\n,(15)\nall theMvectors being taken at x= +0.\nSince the AFM layer thickness is small compared to the spin di ffusion\nlength and the exchange length, we may use the macrospin appr oximation\nwhich was described in detail in Ref. [9]. In this approximat ion, the\nmagnetization changes slowly within the layer thickness. T his allows to\nwrite\n∂JM\n∂x≈JM(LAFM)−JM(+0)\nLAFM=−JM(+0)\nLAFM, (16)\nbecause the magnetization flux is equal to zero at the interfa ce between\nAFM and the nonmagnetic layer closing the electric circuit, JM(LAFM) =\n0. This allows to exclude the terms with space derivative fro m Eq. (9).\nIn the rest terms, M(x, t) andL(x, t) quantities are replaced with their\nvalues at x= 0. Then Eq. (9) takes a more simple form:\n∂M\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nM×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nL×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[M×H]+γ[M×Hd]\n+1\n2γ(β+β′)(M·n)[M×n]+1\n2γ(β−β′)(L·n)[L×n]\n+K/bracketleftBig\nM×/bracketleftBig\nM׈MF/bracketrightBig/bracketrightBig\n+P/bracketleftBig\nM׈MF/bracketrightBig\n= 0, (17)\nwhere\nK=µBQ\neLAFMM2j, P =γαsdµBτQ\neLAFMj. (18)\nThe term with delta function does not present here, since it i s taken into\naccount in the boundary conditions.\nNow we are to use again the macrospin approximation to exclud e the\nspace derivatives from Eq. (10), too.\nOwing to known relationships [21] between MandLvectors, namely,\nM2+L2= 4M2\n0and (M·L) = 0, we have the following conditions:\n/parenleftbigg\nM·∂M\n∂t/parenrightbigg\n+/parenleftbigg\nL·∂L\n∂t/parenrightbigg\n= 0,/parenleftbigg\nL·∂M\n∂t/parenrightbigg\n+/parenleftbigg\nM·∂L\n∂t/parenrightbigg\n= 0.(19)\n6By substituting Eqs. (10) and (17) in (19) we find that conditi ons (19)\nare fulfilled if the terms in (10)\n1\n2γ(α+α′)/bracketleftbigg\nL×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nM×∂2L\n∂x2/bracketrightbigg\n≡X(20)\nsatisfy the following equations:\n(X·M)+K/parenleftBig\nL·/bracketleftBig\nM×/bracketleftBig\nM׈MF/bracketrightBig/bracketrightBig/parenrightBig\n+P/parenleftBig\nL·/bracketleftBig\nM׈MF/bracketrightBig/parenrightBig\n= 0,\n(X·L) = 0. (21)\nLet us decompose the considered Xvector on three mutually orthog-\nonal vectors:\nX=aM+bL+cγ[L×M]. (22)\nThe substitution (22) in (21) gives a=K/parenleftBig\nL·ˆMF/parenrightBig\n−P/parenleftBig\n[L×M]·ˆMF/parenrightBig\n,\nb= 0. As to ccoefficient, it is a current-induced correction to the co-\nefficient of γ[L×M] term in Eq. (10), i.e., a correction to the uniform\nexchange constant Λ. Let us estimate the correction. Multip lying (22)\nscalarly by [ L×M] with (20) taking into account gives\nc=1\nM2L2/parenleftbigg\n[L×M]·/braceleftbigg1\n2γ(α+α′)/bracketleftbigg\nL×∂2M\n∂x2/bracketrightbigg\n+1\n2γ(α−α′)/bracketleftbigg\nM×∂2L\n∂x2/bracketrightbigg/bracerightbigg/parenrightbigg\n=1\n2/braceleftbigg\n(α+α′)1\nM2/parenleftbigg\nM·∂2M\n∂x2/parenrightbigg\n−(α−α′)1\nL2/parenleftbigg\nL·∂2L\n∂x2/parenrightbigg/bracerightbigg\n. (23)\nIt is seen that c∼α/L2\nAFM, while Λ ∼α/a2, where ais the lattice\nconstant [21]. Since LAFM≫a, the mentioned correction to Λ may be\nneglected.\nAs a result, Eq. (10) takes the form\n∂L\n∂t−1\n2κ\nM0/braceleftbigg/bracketleftbigg\nL×∂M\n∂t/bracketrightbigg\n+/bracketleftbigg\nM×∂L\n∂t/bracketrightbigg/bracerightbigg\n+γ[L×H]+γ[L×Hd]−/parenleftbigg\nγΛ−P\nM/parenrightbigg\n[L×M]\n+1\n2γ(β+β′)(M·n)[L×n]+1\n2γ(β−β′)(L·n)[M×n]\n+K/bracketleftBig\nL×/bracketleftBig\nM׈MF/bracketrightBig/bracketrightBig\n−P1\nM2/bracketleftBig\n[L×M]×/bracketleftBig\nM׈MF/bracketrightBig/bracketrightBig\n= 0. (24)\nHere, Λ constant contains also the equilibrium contributio n of the con-\nduction electrons −αsdm/M.\nEquations (17) and (24) are the result of applying the macros pin con-\ncept to AFM. It is shown that such an approximation may be just ified\nformally for AFM layer. Earlier, it was justified for FM layer s [1, 2] and\ngeneralized [9] with spin injection taking into account. Th e macrospin\napproach corresponds well to experimental conditions and s implifies cal-\nculations substantially. The terms with Kcoe���cient in Eqs. (17), (24)\ndescribe effect of STT mechanism, while the terms with Pcoefficient take\nthe spin injection effect into account.\n74 The magnetization wave spectrum and\ndamping\nWe assume that the easy anisotropy axis lies in the plane of AF M layer\nand is directed along yaxis, the FM magnetization vector is parallel to\nthe positive direction of zaxis, the external magnetic field is parallel to z\naxis too (see Fig. 1).\nWe are interesting in behavior of small fluctuations around t he steady\nstateM={0,0,Mz},L={0,Ly,0}, i.e. thesmallquantities Mx, My,/tildewiderMz=\nMz−Mz, Lx,/tildewideLy=Ly−Ly, Lz.\nLet us project Eqs. (17), (24) to the coordinate axes and take the\nterms up to the first order. The zero order terms are present on ly in the\nprojection of Eq. (24) to xaxis. They give\nMz=Hz+P\nγ\nΛ+1\n2(β−β′)≈Hz+P\nγ\nΛ,\nLy=±/radicalBig\n4M2\n0−M2\nz≈ ±2M0. (25)\nNote that the spin-polarized current takes part in creating magnetic\nmoment together with the external magnetic field dueto thesp in injection\ninduced interaction of the electron spins with the lattice [ 22, 23], which\nPparameter in Eq. (25) corresponds to. Such an interaction le ads to\nappearance of an effective magnetic field parallel to the inje ctor magneti-\nzation. As a result, a canted antiferromagnet configuration may be create\nwithout magnetic field. However, such a configuration corres ponds to\nparallel orientation of FM and AFM layers, M/bardblMF. As is shown below,\nthe instability does not occur with this orientation, so tha t an external\nmagnetic field is to be applied to reach instability.\nWith Eq. (25) taking into account, the equations for the first order\nquantities take the form\n∂Mx\n∂t−1\n2κ\nM0/braceleftbigg\n−Mz∂My\n∂t+Ly∂Lz\n∂t/bracerightbigg\n+(γHz+P)My\n−1\n2γ(β+β′)MzMy−1\n2γ(β−β′)LyLz+KMzMx= 0,(26)\n∂My\n∂t−1\n2κ\nM0Mz∂Mx\n∂t−(γHz+P+4πγMz)Mx+KMzMy= 0,(27)\n∂/tildewiderMz\n∂t+1\n2κ\nM0Ly∂Lx\n∂t+1\n2γ(β−β′)LyLx= 0, (28)\n∂Lx\n∂t−1\n2κ\nM0/braceleftBigg\nLy∂/tildewiderMz\n∂t−Mz∂/tildewideLy\n∂t/bracerightBigg\n−γHzLy\nMz/tildewiderMz= 0,(29)\n∂/tildewideLy\n∂t−1\n2κ\nM0Mz∂Lx\n∂t−1\n2γ(β−β′)MzLx= 0, (30)\n8∂Lz\n∂t+1\n2κ\nM0Ly∂Mx\n∂t+(γHz+P+4πγMz)Ly\nMzMx+KMzLz= 0.(31)\nThe set of equations (26)–(31) splits up to two mutually inde pendent\nsets with respect to ( Mx, My, Lz) and (Lx,/tildewideLy,/tildewiderMz). They describe two\nindependent spectral modes, one of them corresponds to prec ession of the\nAFMmagnetization vectoraroundthemagnetic field, while an other tope-\nriodic changes ofthevectorlength alongthemagnetic field. Webeginwith\nthe spectrum and damping of the first mode. We consider monoch romatic\noscillation with ωangular frequency and put Mx, My, Lz∼exp(−iωt).\nThen we obtain from Eqs. (26), (27), (31)\n/parenleftbig\n−iω+KMz/parenrightbig\nMx+/braceleftbigg\nγHz+P−1\n2γ(β+β′)Mz−1\n2iκω\nM0Mz/bracerightbigg\nMy\n−/braceleftbigg1\n2γ(β−β′)−1\n2iκω\nM0/bracerightbigg\nLyLz= 0, (32)\n/parenleftbig\n−iω+KMz/parenrightbig\nMy−/braceleftbigg\nγHz+P+4πγMz−1\n2iκω\nM0Mz/bracerightbigg\nMx= 0,(33)\n/parenleftbig\n−iω+KMz/parenrightbig\nLz+/braceleftbigg\nγ(Λ+4π)+1\n2γ(β−β′)−1\n2iκω\nM0/bracerightbigg\nLyMx= 0.\n(34)\nNote that aforementioned additivity (in the algebraic sens e, the sign\ntakingintoaccount) oftheexternalmagnetic field andthein jection-driven\neffective field takes place not only in the steady magnetizati on (25), but\nalso in the oscillations of the magnetization and antiferro magnetism vec-\ntors, so that both fields appear in Eqs. (32), (33) “on an equal footing”.\nUsually, Λ ≫4π, β, β′. With these inequalities and stationary so-\nlution (25) taking into account we find the dispersion relati on for the\nmagnetization oscillation\n(1+κ2)ω2+2iνω−ω2\n0= 0, (35)\nwhere\nω0=/radicalBig\n2γ2HAHE+(KMz)2+(γHz+P)2, (36)\nν=κγHE+KMz, (37)\nHE= ΛM0is the exchange field, HA= (β−β′)M0is the anisotropy field.\nFormulae (36)and(37)(without currentterms KMzandP)coincide with\nknownones [21,26]. At HE∼106–107G,HA∼103Gwehaveoscillations\nin THz range, ω0∼1012s−1. In absence of current the damping is rather\nhigh: at κ∼10−2\nν\nω0=κ/radicalbigg\nHE\n2HA∼1. (38)\n9Let us consider the contribution of spin-polarized current to the fre-\nquency and damping of AFM resonance. At first we consider STT m ech-\nanism effect [1, 2]. According to (18) and (25),\nKMz=µBQΛ\neLAFMHzj. (39)\nAtHz<0, that corresponds to direction of the magnetic field (and,\ntherefore, the AFM magnetization) opposite to the FM magnet ization,\nthis quantity is negative. The total attenuation becomes ne gative also\n(an instability occurs), if\nj >eκγM 0|Hz|LAFM\nµBQ≡j0. (40)\nAtκ∼10−2,γM0∼1010s−1,|Hz| ∼102G,LAFM∼10−6cm,Q∼1\nwe have j0∼105A/cm2. Atjnear toj0weakly damping THz oscilla-\ntion can be obtained. At j > j 0, instability occurs which may lead to\neither self-sustained oscillations, or a dynamic stationa ry state. The lat-\nter disappears with the current turning off. To answer the que stion about\nfuture of the instability it is necessary to go out the scope o f the linear\napproximation.\nThespin-polarized currentcontributesalsototheoscilla tion frequency.\nAt the mentioned parameter values, we have |KMz| ∼1012s−1that is\ncomparable with the frequency in absence of the current. Thi s allows\ntuning the frequency by the current or excite parametric res onance by\nmeans of the current modulation.\n5 Current-induced spin injection effect\nNow let us discuss the injection mechanism effect [22, 23]. As mentioned\nbefore, the role of the mechanism is reduced to addition of an effective\nfieldP/γto the external magnetic field. At reasonable parameter valu es,\nthat field is much less than the exchange field HE, so that it does not\ninfluence directly the eigenfrequency (36). Nevertheless, that field can\nmodify substantially the contribution of the STT mechanism , because\nEq. (39) with (25) taking into account now takes the form\nKMz=µBQΛ\neLAFM(Hz+P/γ)j. (41)\nSuch a modification leads to substantial consequences. At Hz<0,P <\nγ|Hz|the instability threshold (40) is lowered, since |Hz|−P/γdifference\nappears now instead of |Hz|. If, however, P > γ|Hz|then the AFM\nmagnetization steady state\nMz=Hz+P/γ\nΛ(42)\nbecomes positive that corresponds to the parallel (stable) relative orienta-\ntion of the FM and AFM layers. In this case, the turning on curr ent leads\nto switching the antiparallel configuration (stated before hand by means\n10of an external magnetic field) to parallel one. With turning o ff current,\nthe antiparallel configuration restores.\nSincethementionedinjection-drivenfielddependsonthecu rrent(see(18)),\nthe instability condition (40) is modified and takes the form\nj0\n1+η< j j 0/ηthe antiparallel configuration switches to parallel one. Th e\nrelative contribution of the injection mechanism is determ ined with η\nparameter. At typical values, αsd∼104,κ∼10−2,γM0∼1010s−1,\nτ∼10−12s, this parameter is of the order of unity, so that the injecti on\neffect may lower noticeably the instability threshold.\nNow let us return to the set of equations (26)–(31) and consid er the\nsecond mode describing with Eqs. (28)–(30). The current infl uences this\nmode by changing steady magnetization Mzdue to the injection effective\nfield effect (see (26)), while the STT mechanism does not influe nce this\nmode. A calculation similar to previous one gives the former dispersion\nrelation (35), but now\nω2\n0= 2γ2HEHAγHz\nγHz+P, (44)\nν=κγHEγHz\nγHz+P. (45)\nAtHz<0,P >|Hz|, thatcorrespondstocurrentdensity j > j0/η, the\ntotal attenuation becomes negative, while the frequency be comes imagi-\nnary, that means switching the antiparallel configuration t o parallel one.\nThus, current does not cause instability of that mode.\n6 Easy plane type antiferromagnet\nLetusconsiderbrieflythesituation whereAFMhaseasy-plan eanisotropy.\nWe take the AFM layer yzplane as the easy plane and xaxis as the (hard)\nanisotropy axis. The magnetic field, as before, is directed a longzaxis.\nWithout repeating calculations, similar to previous ones, we present\nthe results. A formal difference appears only in Eq. (36) for t he eigenfre-\nquencyω0of the first of the modes considered above. We have for that\nfrequency\nω0=/radicalBig\n(γHz+P)2+(KMz)2. (46)\nThe damping has the former form (37), so that the instability threshold\nis determined with former formula (43).\nIn absence of the current ( K= 0, P= 0) with not too small damping\ncoefficient κ, the frequency appears to be much less than damping, so\nthat the corresponding oscillations are not observed. The c urrent effect\nincreases the frequency, on the one hand, and decreases the d amping (at\nHz<0), on the other hand, that allows to observe oscillation reg ime.\n117 Fluctuation effect\nIt follows from Eq. (43) that the threshold current density i s proportional\ntotheexternalmagnetic fieldstrength |Hz|anddecreases withthefield. A\nquestion arises about permissible lowest limit of the total field|Hz|+P/γ.\nIn accordance with Eq. (25), such a limit may be the field which create\nmagnetization |Mz|comparable with its equilibrium value due to thermal\nfluctuations. Let us estimate this magnetization and the cor responding\nfield.\nThe AFM energy change in Vvolume under canting the sublattice\nmagnetization vectors with θ <180◦angle between them is\n∆E= ΛM2\n0(1−cosθ)V=1\n2ΛVM2\nz, (47)\nthe anisotropy energy being neglected compared to the excha nge energy.\nThe equilibrium value of the squared magnetization is calcu lated using\nthe Gibbs distribution:\n/an}bracketle{tM2\nz/an}bracketri}ht=∞/integraltext\n−∞M2\nzexp/parenleftbigg\n−ΛVM2\nz\n2kT/parenrightbigg\ndMz\n∞/integraltext\n−∞exp/parenleftbigg\n−ΛVM2\nz\n2kT/parenrightbigg\ndMz=kT\nΛV(48)\n(strictlyspeaking, themagnetization maybechangedwithi n(−2M0,2M0)\ninterval, however, Λ VM2\n0≫kT, so that the integration limits may be\ntaken infinity).\nTo observe the effects described above, the magnetization Mzwhich\nappears under joint action of the external field and the curre nt (see (25))\nshould exceed in magnitude the equilibrium magnetization /an}bracketle{tM2\nz/an}bracketri}ht1/2. At\nthe current density j=j0/(1+η) corresponding to the instability thresh-\nold, this condition is fulfilled at magnetic field\n|Hz|>/radicalbigg\nΛkT\nV(1+η)≡Hmin. (49)\nAt Λ∼104,η∼1,LAFM∼10−6cm and lateral sizes of the switched\nelement 10 ×10µm2we have V∼10−12cm3andHmin≈30 G at room\ntemperature. This limit can be decreased under larger eleme nt size.\nIt should be mentioned also about other mechanisms of AFM can ting.\nThe most known and studied one is the relativistic Dzyaloshi nskii–Moria\neffect(see, e.g.[21,27]). Besides, possible mechanismsha vebeendiscussed\ndue to competition between sdexchange and direct exchange interaction\nof the magnetic ions in the lattice [28]. At the same time, the re are no\nindications, to our knowledge, about measurements of canti ng in conduc-\ntive AFM. So, present theory is related to conductive AFM, in which the\nlattice canting is determined with external magnetic field.\n8 Conclusions\nThe obtained results show a principal possibility of contro lling frequency\nand damping of AMF resonance in FM–AFM junctions by means of s pin-\npolarized current. Under low AFM magnetization induced by a n external\n12magnetic field perpendicular totheantiferromagnetism vec tor, thethresh-\nold current density corresponding to occurring instabilit y is less substan-\ntially than in the FM–FM case. Near the threshold, the AFM res onance\nfrequency increases, while damping decreases, that opens a possibility of\ngenerating oscillations in THz range.\nAcknowledgments\nThe authors are grateful to Prof. G. M. Mikhailov for useful d iscussions.\nThe work was supported by the Russian Foundation for Basic Re -\nsearch, Grant No. 10-02-00030-a.\nReferences\n[1] J.C. Slonczewski. J. Magn. Magn. Mater. 159, L1 (1996).\n[2] L. Berger. Phys. Rev. B 54, 9353 (1996).\n[3] J.A. Katine, F.J. Albert, R.A. Buhrman, E.B. Myers, D.C. Ralph.\nPhys. Rev. Lett. 84, 3149 (2000).\n[4] M. Tsoi, A.J.M. Jansen, J. Bass, W.-C. Chiang, M. Seck, V. Tsoi, P.\nWyder Phys. Rev. Lett. 80, 4281 (1998).\n[5] A. Yamaguchi, T. Ono, S. Nasu, K. Miyake, K. Mibu, T. Shinj o.\nPhys. Rev. Lett. 92, 077205 (2004).\n[6] J.C. Sankey, P.M. Braganca, A.G.F. Garcia I.N. Krivorot ov, R.A.\nBuhrman, D.C. Ralph. Phys. Rev. Lett. 96, 227601 (2006).\n[7] M. Watanabe, J. Okabayashi, H. Toyao, T. Yamaguchi, J. Yo shino.\nAppl. Phys. Lett. 92, 082506 (2008).\n[8] Yu.V. Gulyaev, P.E. Zilberman, A.I. Krikunov, E.M. Epsh tein.\nTechn. Phys. 52, 1169 (2007).\n[9] Yu.V. Gulyaev, P.E. Zilberman, A.I. Panas, E.M. Epshtei n. J. Exp.\nTheor. Phys. 107, 1027 (2008).\n[10] Yu.V. Gulyaev, P.E. Zilberman, S.G. Chigarev, E.M. Eps htein.\nTechn. Phys. Lett. 37, 154 (2011).\n[11] Yu.V.Gulyaev, P.E.Zilberman, E.M. Epshtein.J.Commu n.Technol.\nElectron. 56, 863 (2011).\n[12] V.K.Sankaranarayanan, S.M.Yoon, D.Y.Kim, C.O. Kim, C .G. Kim.\nJ. Appl. Phys. 96, 7428 (2004).\n[13] A. S. N´ u˜ nez, R. A. Duine, P. Haney, A.H. MacDonald. Phy s. Rev. B\n73, 214426 (2006).\n[14] Z. Wei, A. Sharma, A.S. Nunez, P.M. Haney, R.A. Duine, J. Bass,\nA.H. MacDonald, M. Tsoi. Phys. Rev. Lett. 98, 116603 (2007).\n[15] Z. Wei, A. Sharma, J. Bass, M. Tsoi. J. Appl. Phys. 105, 07D113\n(2009).\n[16] J. Basset, Z. Wei, M. Tsoi. IEEE Trans. Magn. 46, 1770 (2010).\n13[17] S. Urazhdin, N. Anthony. Phys. Rev. Lett. 99, 046602 (2007).\n[18] H.V. Gomonay, V.M. Loktev. Low Temp. Phys. 34, 198 (2008).\n[19] H.V. Gomonay, V.M. Loktev. Phys. Rev. B 81, 144127 (2010).\n[20] K.M.D. Hals, Y. Tserkovnyak, A. Brataas. Phenomenolog y of\ncurrent-induced dynamics in antiferromagnets, arXiv:101 2.5655v1\n[cond-mat.mes-hall].\n[21] A.I. Akhiezer, V.G. Baryakhtar, S.V. Peletminslii. Sp in Waves,\nNorth-Holland Pub. Co., Amsterdam, 1968.\n[22] C. Heide, P.E. Zilberman, R.J. Elliott, Phys. Rev. B 63, 064424\n(2001).\n[23] Yu.V. Gulyaev, P.E. Zilberman, E.M. Epshtein, R.J. Ell iott, JETP\nLett.76, 155 (2002).\n[24] Yu.V. Gulyaev, P.E. Zilberman, A.I. Panas, E.M. Epshte in.\nPhysics — Uspekhi 52, 335 (2009).\n[25] Yu.V. Gulyaev, P.E. Zilberman, E.M. Epshtein, R.J. Ell iott. J. Exp.\nTheor. Phys. 100, 1005 (2005).\n[26] A.G.Gurevich, G.A.Melkov.Magnetizsation Oscillati ons andWaves,\nCRC Press, Boca Raton, FL, 1996.\n[27] V.E. Dmitrienko, E.N. Ovchinnikova, J. Kokubun, K. Ish ida. JETP\nLett.92, 383 (2010).\n[28] J.M. Robinson, P. Erd¨ os. Phys. Rev. B 6, 3337 (1972).\n14" }, { "title": "2009.01057v1.Effect_of_the_Optical_Pumping_and_Magnetic_Field_on_the_States_of_Phase_Separation_Domains_in_Eu__0_8_Cr__0_2_Mn_2O_5.pdf", "content": "1 \n \n \nEffect of the Optical Pumping and Magnetic Field on the States of Phase \nSeparation Domains in Eu 0.8Ce0.2Mn 2O5 \n \nE. I. Golovenchitsa, B. Kh. Khannanova, and V. A. Saninaa,* \n \na Ioffe Institute, St. Petersburg, 194021 Russia \n*e-mail: sanina@mail.ioffe.ru \n \nThe effect of optical pumping and applied magnetic field on the characteristics of ferromagnetic \nlayers in one-dimensional superlattices is studied. At low enough temperatures, these layers \ncorrespond to phase separation domains in RMn 2O5 and R 0.8Ce0.2Mn 2O5 multiferroics. The \nformation of such domains occurs owing to the charge ordering of Mn3+ and Mn4+ ions and to the \nfinite probability for eg electrons to tunnel between these pairs of ions. The volume occupied by \nsuch superlattices is rather small, and they can be treated as isolated ferromagnetic semiconductor \nheterostructures, spontaneously formed in the host crystal. The sequences of ferromagnetic \nresonances related to the s uperlattice layers in Eu 0.8Ce0.2Mn 2O5 are studied. The characteristics of \nthese resonances give information on the properties of such layers. For the first time, it is \ndemonstrated that the optical pumping gives rise to a new metastable state of superlattices, which \ncan be recovered by the magnetic field cycling to the state existing before the optical pumping. It \nis found that the superlattices recovered by the magnetic field exist up to temperatures higher than \nthose in as -grown crystals. \n \n \n The presence of equal numbers of Mn3+ and Mn4+ manganese ions is a characteristic feature of \nRMn 2O5 multiferroic compounds ( Pbam symmetry), ensuring the formation of charge ordering. \nMn4+ ions are surrounded by oxygen octahedra and located in the layers having the positions z = \n0.25c and (1 − z) = 0.75 c. These ions contain three t2g electrons and have an empty degenerate \norbital doublet corresponding to the eg state. Mn3+ ions contain three t2g electrons in their 3d shell \nand one eg electron in the orbital doublet. These ions are located in the layers having the positions \nz = 0.5 c, and their local environment in the form of pentagonal pyramids has no center of \nsymmetry. R3+ ions with the environment similar to that of Mn3+ are located in the z = 0 layers [1]. \nCharge ordering and the finite probability of the eg electron transfer between neighboring Mn3+–\nMn4+ ion pairs (double exchange [2, 3]) are key factors determining the multiferroic characteristics \nof RMn 2O5 compounds in the range from low to room temperatures. Magnetic ordering with a \ncomplicated structure occurs below the Neel temperature TN ≈ 40−45 K, whereas the ferroelectric 2 \n ordering induced by the magnetic order arises below the Curie temperature TC ≈ 35−40 K [4]. The \npolar order is mainly related to the exchange striction induced by neighboring Mn3+ and Mn4+ ion \npairs located along the b axis and having alternating ferromagnetic and antiferromagnetic spin \norienta tions [5]. The eg electron transfer between Mn3+ and Mn4+ ions located in adjacent layers \nperpendicular to the c axis leads to the formation of local polar regions of phase separation with a \ndifferent distribution of Mn3+ and Mn4+ ions as compared to that in the initial host crystal. The \nstates resulting from phase separation were studied in detail in pure RMn 2O5 multiferroic \nmanganites, as well as in the doped ones, R0.8Ce0.2Mn 2O5 (R = Eu, Gd, Bi, Tb, and Er), having the \nsame Pbam symmetry [6 –13]. In these materials, the comparative studies of dielectric and \nmagnetic properties, heat capacity, X -ray diffraction, Raman light scattering [6,7], electric \npolarization [8 –11], and μ-SR [12, 13] were carried out. \n Phase separation was observed in undoped RMn 2O5 crystals, whereas their doping with Ce4+ \nions substituting R3+ ions leads to a significant increase in the density of neighboring Mn3+–Mn4+ \nion pairs in the planes perpendicular to the c axis. In R0.8Ce0.2Mn 2O5 and in the z = 0 plane, the \nR3+ = Ce4+ + e reaction provides additional electrons. These electrons transform Mn4+ ions to Mn3+ \nones in the z = 0.25c and 1 − z = 0.75 c planes. As a result, the number of Mn3+–Mn4+ pairs and the \nnumber of phase separation domains increase. At the same time, the phas e separation domains \noccupy a small part of the crystals both in RMn 2O5 and in R0.8Ce0.2Mn 2O5 [6, 7]. Phase separation \ndomains are formed in RMn 2O5 and in R0.8Ce0.2Mn 2O5 in much the same way as those in \nLaAMnO 3 (A = Sr, Ca, and Ba) compounds, which also contain Mn3+ and Mn4+ ions [3, 14 –16]. \nThese domains are formed because of the balance between strong interactions occurring in the \nsubsystem of Mn ions, such as the double exchange with the characteristic energy of about 0.3 eV, \nthe Jahn –Teller interactio n (0.7 eV), and the Coulomb repulsion (1 eV). That is why these domains \nin RMn 2O5 and in R0.8Ce0.2Mn 2O5 exist in a wide temperature range from low temperatures to 3 \n Fig. 1. (a) Schematic image of 1D ferromagnetic superlattices (shaded areas) located in the bulk \nof host crystal (white background); magnetic field H = 0. (b) Schematic image of one of such \nsuperlattices consisting of L N ferromagnetic layers perpendicular to the c axis having different \ndensities of Mn3+−Mn4+ ion pairs and of eg electrons located in potential wells of different depths \n(shaded areas corresponding to different energies E). In Eu 0.8Ce0.2Mn 2O5, the linear size of \ndomains with superlattices is about 700 Å [6, 7]. \n \nthose above room temperature. At T < 60 K, phase separation domains are isolated one -\ndimensional (1D) superlattices consisting of ferromagnetic layers containing Mn3+ and Mn4+ ions \nwith different relative densities, as well as eg electrons changing their charge (Fig. 1). These \ndomains manifest themselves as a serie s of ferromagnetic resonances (FMR) related to the \nindividual layers of the superlattices. The characteristic features of such resonances reveal the \nproperties of these layers and of the superlattice as a whole [17–20]. \n The aim of this work is to study the effect of optical pumping, applied magnetic field, and \ntemperature on the properties of 1D superlattices in Eu0.8Ce0.2Mn2O5 single crystals. Optical \npumping can affect the electron density in the layers of superlattices and, thus, result in its \nredistribution in these layers. This, in turn, should change the ratio of the numbers of ions with \ndifferent valences in the layers, which can be detected by changes in the characteristics of the FMR \nlines. \n4 \n Eu0.8Ce0.2Mn 2O5 single crystals were grown b y spontaneous crystallization from the solution –\nmelt [21, 22]. They were platelets 1 −3 mm thick with an area of 3−5 mm2. For FMR measurements, \nwe used a transmission type magnetic resonance spectrometer with a low -amplitude magnetic \nmodulation. The measurements were performed in the temperature range of 13 –300 K at \nfrequencies of 30 –40 GHz in an applied magnetic field up to 2 T generated by an electromagnet. \nThe cryostat with optical windows was located in the microwave channel ensuring a uniform \ndistribution of the microwave field close to the sample. The microwave radiation (with the wave \nvector k) was directed along the c axis perpendicular to the platelet plane. The static magnetic field \nH was oriented along the a axis of the crystal and perpendicular to the direction of the microwave \nfield h. The measured FMR signals were amplified by an SR530 lock -in amplifier. Naturally \nfaceted single crystals were used. The symmetry of the crystals and their composition were \ndetermined by the X -ray pha se analysis and X -ray fluorescence technique, respectively. Optical \npumping was performed by a solid -state pulsed neodymium laser LTIPCH -8 with the simultaneous \ngeneration of the first (1.06 μm) and second (532 nm) harmonics. The second harmonic \ncorrespond s to the 7F0−5D1 transition in Eu3+ ions. The 5D1 state is located within the electron –\nphonon band of Mn ions in RMn 2O5, with the edge beginning at about 485 nm [23]. \n Earlier, the FMR studies of layers in 1D superlattices in a series of RMn 2O5 (R = Eu, Er, Tb, \nand Gd) and R 1 – x CexMn 2O5 (R = Eu and Gd , x = 0.2, 0.25 ) crystals demonstrated that the magnetic \nfields at which individual FMR lines are observed differ only slightly and are independent of the \nR ion type [17 –20]. In this case, Er3+ and Tb3+ (unlike Gd3+) are strongly coupled to the lattice. \nThis means that the FMR resonance fields are determined by the internal fields far exceeding the \nanisotropy fields characteristic of the superlattice layers; the latter fields should depend on specific \nR ions. Indeed, as mentioned above, the equilibrium state of 1D layers in the superlattices is \ndetermined by the balance of interactions just within the subsystem of Mn ions. In this case, the \nlayers in the 1D superlattices can be treated as isotropic ferr omagnetic layers with ferromagnetic 5 \n boundaries between them, which do not put obstacles in the way of the eg electron transfer between \nthe layers related to the double exchange (Fig. 1b) [17–20]. A model describing the spin -wave \nexcitations in superlattice layers was developed in these works. The model is based on the \ndispersion equation for spin waves in isotropic ferromagnetic films forming 2D multilayers [24]. \nIt is also demonstrated that the FMR lines in 1D superlattices are observed in individual frequ ency \nbands separated by the ranges in which the FMR is not observed. This occurs because the 1D \nsuperlattices (similar to semiconductor superlattices [25]) have a band structure consisting of \nminibands separated by gaps. The frequency of 34.5 GHz, at which the measurements were \nperformed, is located inside one of the main minibands in Eu 0.8Ce0.2 Mn 2O5 (29–36 GHz) [17]. \n It was reported in [17, 18] that the formation of the equilibrium state of superlattices for \nEuMn 2O5 and Eu0.8Ce0.2 Mn 2O5, in which a set of FMR lines was observed, required the magnetic \ncycling (subsequently increasing and decreasing the magnetic field) of as -grown samples. The \ndynamically equilibrium (ground) state of the superlattices is achieved after triple magnetic \ncycling in the r ange of 0 –20 kOe. Note that the ground state of the superlattices was recovered \nupon the cooling of the sample to helium temperatures after its heating to room temperature and \nits long (several weeks) keeping at this temperature. The repeated cycling of the magnetic field is \nnot necessary. \n For the 1 -mm-thick Eu0.8Ce0.2 Mn 2O5 sample with an area of 5 mm2, the sequence of FMR \nlines in the equilibrium state is shown in Fig. 2a. Generally speaking, these lines correspond to \nthose measured earlier in [17, 18]. We can see five FMR lines originating from different layers of \nsuperlattices schematically shown in Fig. 1b . The most intense central L0 line appears in the \napplied magnetic field corresponding to the FMR at this frequency for an isotropic fe rromagnet \nwith the g factor g = 2. Less intense lines, to the left of the main resonance (L1 and L2) and to the \nright of it (R1 and R2), are approximately symmetrically located in lower and higher magnetic \nfields with respect to that corresponding to the L0 line. In this case, the most intense lines (L1, L0, \nand R1) are doubled. At somewhat lower fields, less intense wide lines are observed (Fig. 2a). 6 \n \nFig. 2. Distribution of intensities for the observed FMR lines in Eu 0.8 Ce0.2 Mn 2O5 at 34.5 GHz. \nThe sensitivity of the amplifier is 5 μV; H || a . The magnetic field was varied at a rate of 1.2 \nkOe/min. \n(a) As -grown sample after the third cycle of the magnetic field increase (dynamically equilibrium \nstate of a 1D superlattice); T = 20 K. \n(b) Dynamically equilibrium state of the superlattice (black curve) is compared to the distribution \nof intensities for the FMR lines after optical pumping for 1 min by 15 ns pulses with a power of \n0.5 MW and a repetition rate of 10 Hz. Magnetic field H decreases (red curve); T = 22 K. \n(c) Changes in the state of the sample after 16 h of its natural heating up to 250 K and its subsequent \ncooling down to 17.5 K, with an increase in H (black curve) and with its subsequent decrease \n(red curve). \n(d) Recovery of the state similar to the initial dynamically equilibrium one (Fig. 2a) during the \nthird cycle of the magnetic field growth; T = 17.5 K. \n \n \n7 \n In [17, 18], the FMR lines were observed at temperatures below 60 K. Their intensity \ndecreases gradually with an increase in the temperature. The resonant magnetic fields for a \nsequence of FMR lines exhibit a linear frequency dependence ωn = γn(H + Hneff). Here, ωn are the \ncircular frequencies for n lines ( n = L2, L1, L0, R1, and R2), γn are the gyromagnetic ratios, and \nHneff are the internal effective fields responsible for the gaps in ωn(H). The values Hneff are positive \nfor lines L1 and L2 and increase with the number L. For lines R1and R2, the values Hneff are \nnegative. The g factors for the lines R1and R2 are slightly smaller than 2, whereas the g factors \nfor L1 and L2 lines exceed 2 [17, 18]. \n Following the approach reported in [18 –20], we analyze the properties of superlattice layers \nusing the model of semiconductor heterostructures c ontaining alternating L, 0, and R layers, in \nwhich Mn4+ and Mn3+ ions are treated as acceptor and donor impurities, respectively. Analyzing \nthe characteristics of the observed FMR lines related to the layers of superlattices, we revealed the \nproperties of these layers. We assume that the L0 layers contain equal numbers of Mn4+ and Mn3+ \nions; i.e., these layers are fully compensated semiconductors. In such layers, the Fermi level is \nlocated in the center of the band gap, and the insulating state arises in th e L0 layers [26]. The \nabsence of free electrons in the L0 layers explains the maximum intensity of the L0 lines in the \nequilibrium states of superlattices (Fig. 2a). The L and R layers are partially compensated \nsemiconductors: Mn4+ ions play the dominant role in the L layers, whereas the Jahn–Teller Mn3+ \nions responsible for local structural distortions prevail in the R layers. The R layers exhibit \nthe deepest potential wells with the highest electron density (Fig. 1b). According to [17 –19], the \nR layers are characterized by negative Hneff values and by g factors below 2. As mentioned above, \nfor L layers the values Hneff are positive and the corresponding g factors slightly exceed 2. The \ndynamic equilibrium resulting from the balance of competing interactions (double exchange and \nthe Jahn –Teller effect, increasing the electron density in the layers of the superlattices, and the \nCoulomb repulsion of electrons), as well as the charge neutrality of the superlattice as a whole, \nsuggests that the states of the layers should be mutu ally correlated and form a periodic (L−O−R) 8 \n array. That is why the resonance fields of individual FMR lines turn out to be strictly fixed and \nalmost the same in RMn 2O5 and R0.8Ce0.2Mn 2O5 compounds with different R ions [18–20]. \n Let us now discuss the effect of optical pumping on the state of layers in the superlattices. \nPumping was carried out for 1 min at T = 22 K by 15 -ns pulses with a peak power of about 0.5 \nMW and a repetition rate of 10 Hz. In Fig. 2b, we show a sequence of FMR lines (red poi nts on \nthe curve) measured at the same temperature a few minutes after the end of optical pumping. These \nlines arise with an increase in the magnetic field at a rate of 1.2 kOe/min. These data are plotted \nagainst the background of the equilibrium state of the layers in the superlattice (black points and \nthe curve). The sequence of FMR lines is observed at the same resonance fields as that before \npumping, but the intensities of the L0 and R1 lines change. The intensities of other lines remain \nalmost unchange d. Before pumping, the L0 line is twice as intense as the R1 line (Fig. 2a), whereas \nthe intensities of these lines after pumping become equal (Fig. 2b). We suppose that, upon \nrelaxation of optical excitations in the electron -phonon band related to Mn ions , electrons \nappearing in the layers L0 and R1 of superlattices turn out to be in excess as compared to the initial \nelectron density formed because of doping with Ce4+ ions. Most of these electrons are localized at \nMn4+ ions (Mn4+ + e = Mn3+), transforming them to Jahn –Teller ions, which lead to local lattice \ndistortions in the corresponding layers, thus reducing the energy of superlattices. An increase in \nthe density of Jahn –Teller Mn3+ ions in the L0 layers results in the deepening of potential wells \ninsid e these layers, bringing their states closer to the R1 states. This suggests an intense exchange \nof electrons between the L0 and R1 layers at T = 22 K, which makes electron densities in these \nlayers equal (Fig. 2b). A new state of superlattices arises, whi ch is long -lived at low temperatures, \nsince it does not disappear for at least 20 min after the end of pumping. \n In Fig. 2c, we demonstrate a sequence of FMR lines for the same sample after 16 h of natural \nheating up to 250 K (after optical pumping) and its subsequent cooling down to 17.5 K (without \nnew optical pumping). Black and red points and lines correspond to increasing and decreasing \nmagnetic fields, respectively. We can see that the intensities of the L2 and L1 lines still do not 9 \n change at inc reasing and decreasing fields. At the same time, as the field increases, the intensities \nof the R1 and R2 lines increase steeply and exceed the intensity of the L0 line. Thus, the state \nforming after optical pumping is not recovered when the sample is cool ed down to 17.5 K after its \nheating and subsequent cooling. This means that the state arising after optical pumping with the \nsame populations of the L0 and R1 layers (Fig. 2b) is long -lived. \n It is natural to assume that the heating of the sample after pumping leads to a certain \nredistribution in the L0, R1, and R2 layers of excess electrons generated by optical pumping, which \nleads to the reduction of the superlattice energy. The heating of the sample should result in an \nincrease in the kinetic en ergy of electrons, thus causing their redistribution among the layers owing \nnot only to tunneling between the L0 and R1 layers (Fig. 2b) but also to the hopping conductivity, \nwhich favors overcoming higher barriers at the boundaries of other layers of supe rlattices. The \npopulation of the deeper potential wells with the maximum number of Mn3+ ions is the most \nprobable. This is the case for the R2 layer, which is nearly empty at low temperatures (Figs. 2a, \n2b). At the same time, two other layers, R1 and L0, a re populated but with a lower probability. \nStructural distortions also increase in them owing to an increase in the number of Jahn –Teller Mn3+ \nions, for which the potential wells in these layers are deeper. In this case, all excess electrons are \nlocalized at Mn3+ ions. As a result, intense narrow FMR lines related to the R2, R1, and L0 layers \nare observed (Fig. 2c, black points and curves). With a subsequent decrease in the magnetic field, \nthe intensities of the L0, R1, and R2 lines decrease steeply (Fig. 2 c, red curves). In this case, the \nR2 line intensity becomes nearly equal to its value in dynamical equilibrium before pumping, \nwhereas the R1 line intensity exceeds that of the L0 line, although only slightly. Only an additional \nincrease in the magnetic field recovers a state similar to that corresponding to the initial \nequilibrium state of the superlattice before pumping (Fig. 2d), but with slightly lower intensities \nof the FMR lines. The cycling of the magnetic field, recovering the ferromagnetic orient ation of \nthe layers in superlattices, enhances the double exchange, leading to the tunneling of electrons 10 \n between the layers in superlattices. It turns out that the resulting state exists up to a higher \ntemperature than the equilibrium state before optic al pumping (Fig. 3). \nFig. 3. (a) Sequence of FMR lines corresponding to the layers in superlattices after optical pumping \nand cycling in the applied magnetic field at various temperatures. Actually coinciding zeros of \nintensity are displaced relative to each other. \n(b) Temperature dependence of intensities for the L0 and R1 lines. \n In Fig. 3, we show the set of the FMR lines in the recovered equilibrium state after optical \npumping at various temperatures (Fig. 3a) and the temperature dependence of the intensities of the \nL0 and R1 lines in this state (Fig. 3b). We can see that the equilibrium states of superlattices after \noptical pumping exist up to T ≈ 115 K, whereas before pumping, they are observed up to T ≈ 60 K \n[12]. \n As mentioned above, more intense and narrower L1, L0, and R1 lines in Fig. 2 are doubled: \nthe low intensity wide satellites are observed near these lines at somehow lower fields. This occurs \nbecause the initial crystal contains not only Ce4+ ions but also Ce3+ ions, although in a smaller \nnumber, which induce local regions with strong lattice distortions because of the presence of alone \n6s2 electron pairs in their outer shells. We previously observed this effect in R0.8Ce0.2Mn 2O5 (R = \nEr and Tb) [27]. In these regions with a ce rtain number of localized electrons, superlattices with \nslightly different parameters are formed. \n11 \n To summarize, it has been demonstrated for the first time that the states of 1D superlattices \n(ferromagnetic semiconductor heterostructures) in Eu 0.8Ce0.2Mn 2O5 multiferroics, which at \nsufficiently low temperatures are phase separation domains, can be controlled by optical pumping \nand applied magnetic field. Optical pumping leads to the formation of a new long -lived state of \nsuperlattices with a different distribution of Mn3+ and Mn4+ ions and of electrons changing their \ncharge in these layers. The cycling of the applied magnetic field in this new state of superlattices \nafter pumping recovers the state with the distribution of intensities of resonance lines similar to \nthat existing before pumping. In this case, the intensities in the sequence of resonance lines \nrecorded after pumping are slightly lower than those before pumping. This is due to the increased \ndensities of electrons and Mn3+ and Mn4+ ion p airs in the layers. However, these lines exist up to \na higher temperature, which means that the new equilibrium state recovered by the magnetic field \nafter optical pumping is more favorable in energy than the state before pumping. \n \nFUNDING \nThis work was supported by the Russian Foundation for Basic Research (project no. 18 -32-\n00241) and by the Presidium of the Russian Academy of Sciences (program 1.4 “Topical \nProblems of Low -Temperature Physics”). \n \nREFERENCES \n \n1. P. G. Radaelli and L. C. Chapon, J. Phys.: Condens. Matter 20, 434213 (2008). \n2. P. G. de Gennes, Phys. Rev. 118, 141 (1960). \n3. L. P. Gor’kov, Phys. Usp. 41, 581 (1998). \n4. Y. Noda, H. Kimura, M. Fukunaga, S. Kobayashi,I. Kagomiya, and K. Kohn, J. Phys.: \nCondens. Matter 20, 434206 (2008). \n5. J. v an den Brink and D. I. Khomskii, J. Phys.: Condens. Matter 20, 434217 (2008). \n6. V. A. Sanina, E. I. Golovenchits, V. G. Zalesskii, S. G. Lushnikov, M. P. Scheglov, S. N. \nGvasaliya, A. Savvinov, R. S. Katiyar, H. Kawaji, and T. Atake, Phys. Rev. B 80, 2244 01 \n(2009). \n7. V. A. Sanina, E. I. Golovenchits, V. G. Zalesskii, and M. P. Scheglov, J. Phys.: Condens. \nMatter 23, 456003 (2011). \n8. V. A. Sanina, E. I. Golovenchits, B. Kh. Khannanov, M. P. Scheglov, and V. G. Zalesskii, \nJETP Lett. 100, 407 (2014). \n9. B. Kh. Khannanov, V. A. Sanina, E. I. Golovenchits, and M. P. Scheglov, JETP Lett. 103, 248 \n(2016). \n10. B. Kh. Khannanov, V. A. Sanina, and E. I. Golovenchits, J. Phys.: Conf. Ser. 572, 012046 \n(2014). 12 \n 11. B. Kh. Khannanov, V. A. Sanina, E. I. Golovenchits, and M. P. Scheglov, J. Magn. Magn. \nMater. 421, 326 (2017). \n12. D. S. Andrievskii, S. I. Vorob’ev, A. L. Getalov, E. I. Golovenchits, E. N. Komarov, S. A. \nKotov, V. A. Sanina, and G. V. Shcherbakov, JETP Lett. 106, 295 (2017). \n13. S. I. Vorob’ev, A. L. Getalo v, E. I. Golovenchits, E. N. Komarov, S. A. Kotov, V. A. Sanina, \nand G. V. Shcherbakov, JETP Lett. 110, 133 (2019). \n14. M. Yu. Kagan and K. I. Kugel’, Phys. Usp. 44, 553 (2001). \n15. J. Lorenzana, J. C. Castellani, and C. di Castro, Europhys. Lett. 57, 704 (2002). \n16. K. I. Kugel, A. L. Rakhmanov, A. O. Sboychakov, F. V. Kustmarsev, N. Poccia, and A. \nBianconi, Supercond. Sci. Technol. 22, 014007 (2009). \n17. E. I. Golovenchits, V. A. Sanina, and V. G. Zalesskii, JETP Lett. 95, 386 (2012). \n18. V. A. Sanina, E. I. Golovenchits, and V. G. Zalesskii, J. Phys.: Condens. Matter 24, 346002 \n(2012). \n19. V. A. Sanina, B. Kh. Khannanov, and E. I. Golovenchits, Phys. Solid State 59, 1952 (2017). \n20. V. A. Sanina, E. I. Golovenchits, V. G. Zalesskii, and B. Kh. Khannanov, J. Phys.: Condens. \nMatter 25, 336001 (2013). \n21. V. A. Sanina, L. M. Sapozhnikova, E. I. Golovenchits, and N. V. Morozov, Sov. Phys. Solid \nState 30, 1736 (1988). \n22. A. V. Babinskii, E. I. Golovenchits, N. V. Morozov, and L. M. Sapozhnikova, Sov. Phys. \nSolid State 34, 56 (1992). \n23. A. S. Moskvin and R. V. Pisarev, J. Low Temp. Phys. 36, 489 (2010). \n24. A. G. Gurevich and G. A. Melkov, Magnetization Oscillations and Waves (Fizmatlit, \nMoscow, 1994; CRC, New York, 1996). \n25. A. P. Silin, Sov. Phys. Usp. 28, 972 (1985). \n26. B. I. Shklovskii and A. L. Efros, Electronic Properties of Doped Semiconductors (Springer, \nHeidelberg, 1984). \n27. B. Kh. Khannanov, E. I. Golovenchits, and V. A. Sanina, Phys. Solid State 62, 660 (2020). " }, { "title": "2008.05268v1.Accelerating_the_laser_induced_demagnetization_of_a_ferromagnetic_film_by_antiferromagnetic_order_in_an_adjacent_layer.pdf", "content": "Accelerating the laser-induced demagnetization of a\nferromagnetic \flm by antiferromagnetic order in an adjacent layer\nI. Kumberg,1E. Golias,1N. Pontius,2R. Hosseinifar,1K. Frischmuth,1I. Gelen,1T.\nShinwari,1S. Thakur,1C. Sch u\u0019ler-Langeheine,2P. M. Oppeneer,3and W. Kuch1\n1Institut f ur Experimentalphysik, Freie Universit at Berlin,\nArnimallee 14, 14195 Berlin, Germany\n2Helmholtz-Zentrum Berlin f ur Materialien und Energie,\nAlbert-Einstein-Stra\u0019e 15, 12489 Berlin, Germany\n3Department of Physics and Astronomy,\nUppsala University, Box 516, SE-75120, Uppsala, Sweden\n(Dated: August 13, 2020)\nAbstract\nWe study the ultrafast demagnetization of Ni/NiMn and Co/NiMn ferromag-\nnetic/antiferromagnetic bilayer systems after excitation by a laser pulse. We probe the\nferromagnetic order of Ni and Co using magnetic circular dichroism in time-resolved pump{probe\nresonant X-ray re\rectivity. Tuning the sample temperature across the antiferromagnetic ordering\ntemperature of the NiMn layer allows to investigate e\u000bects induced by the magnetic order of the\nlatter. The presence of antiferromagnetic order in NiMn speeds up the demagnetization of the\nferromagnetic layer, which is attributed to bidirectional laser-induced superdi\u000busive spin currents\nbetween the ferromagnetic and the antiferromagnetic layer.\n1arXiv:2008.05268v1 [cond-mat.mes-hall] 12 Aug 2020I. INTRODUCTION\nMagnetic recording and storage media will face major challenges in the near future re-\ngarding energy e\u000eciency and data transfer speed due to increased demand and continuing\ngrowth of data storage volume. Much could be gained by switching from \feld-induced\nmanipulation of magnetic order, where speed is limited by spin precession dynamics1{3, to\nmanipulation by ultra short light pulses4{11. To explain ultrafast magnetization dynamics\nfollowing the excitation by an ultrashort laser pulse, a number of di\u000berent approaches are\nbeing discussed. One can distinguish between local mechanisms such as scattering with\n(quasi-)particles5,6,12{14and non-local mechanisms such as transport by superdi\u000busive spin\ncurrents15,16. In heterostructures, the latter may dominate the observed dynamic behavior\nsince they allow transfer of angular momentum between di\u000berent ferromagnetic layers, which\nmay even lead to an entirely di\u000berent dynamic behavior as compared to isolated layers17,18.\nTransient superdi\u000busive spin currents are also expected to occur at the interface between\nan antiferromagnet (AFM) and a ferromagnet (FM). Although unpolarized in the interior\nof an AFM layer, superdi\u000busive currents excited by a laser pulse become spin-polarized in\nthe vicinity of the interface to an FM layer because of spin-polarized interface re\rection and\ntransmission. In this work, we present a study on the ultrafast optical demagnetization of\nan FM/AFM layered system. We show that the presence of antiferromagnetic order in an\nadjacent layer accelerates the demagnetization of the FM layer, which we attribute to the\nexchange of superdi\u000busive spin currents between the two layers.\nFM/AFM layered systems are well known in device design, as the exchange interaction\nbetween the two layers leads to an exchange bias (EB) e\u000bect, resulting in the magnetic\npinning of the FM layer19,20. NiFe/NiO as an FM/AFM system has been already studied\nby time-resolved magneto-optical Kerr e\u000bect in the pioneering works of Ju et al. where an\nunpinning of the interface spins within times \u00141 ps was reported21{23., More recently, the\ntorque on the FM spins in an FM/AFM bilayer after laser excitation was investigated24,\nsuggesting, like the works of Ju et al., that the magnetization in such a system is switchable\nby a laser pulse. In extension of these previous investigations, which were mainly targeting\nthe evolution of EB after laser excitation and the correlated precessional motion of the FM\nmacrospin, we investigate here the ultrafast processes in an FM/AFM bilayer system after an\noptical excitation pulse. Using Ni and Co as ferromagnets and NiMn as an antiferromagnet,\nwe have grown epitaxial FM/AFM bilayer samples with easily accessible N\u0013 eel temperature\n2TNof the AFM layer. We employ resonant X-ray magnetic circular dichroism in re\rectiv-\nity to probe the magnetization with elemental resolution. Comparing the demagnetization\nbehavior above and below TNallows us then to investigate the in\ruence of the magnetic\norder in the NiMn layer on the demagnetization of Ni and Co following the excitation with\na femtosecond laser pulse. We observe a signi\fcantly faster demagnetization for the same\namount of demagnetization of the FM layer if the adjacent NiMn layer is antiferromagneti-\ncally ordered. We discuss this in terms of the exchange of angular momentum through the\nFM/AFM interface by superdi\u000busive spin currents.\nII. EXPERIMENT\nThe samples are grown by molecular beam epitaxy in our home lab at the Freie Uni-\nversit at Berlin. All \flms are prepared on Cu(001) single crystals, which were cleaned by\nmultiple Ar+sputtering and annealing cycles. The surface integrity is veri\fed by LEED\nand the sample's cleanliness by Auger electron spectroscopy. The materials are thermally\nevaporated by electron bombardment from a rod of 99.998% purity for Ni and Co, and from\nMn \rakes with 99.98% purity in a Ta crucible. During evaporation the sample is kept at\nroom temperature and deposition rates of about 1 monolayer (ML) per minute are used. The\npressure is kept below 1 \u000210\u00009mbar while evaporating, with a base pressure of \u00196\u000210\u000010\nmbar. Deposition is monitored by medium-energy electron di\u000braction to ensure layer-by-\nlayer growth. Composition and thickness are additionally checked and calibrated by the\nsignal intensity in Auger electron spectroscopy. After preparation, the samples are capped\nby 20 ML Cu with purity of 99.99%, evaporated from a Ta crucible, to prevent sample oxi-\ndation during transport. We prepared a 20 ML Cu /15 ML Co /20 ML Ni 31Mn69/Cu(001)\nand a 20 ML Cu /12 ML Ni /14 ML Ni 38Mn62/Cu(001) sample. For the AFM layer we chose\nNiMn, an alloy which orders antiferromagnetically around equiatomic concentrations. Fur-\nthermore, EB can occur both in in-plane and out-of-plane geometries25,26and a vanishing of\ndirect exchange coupling between out-of-plane-magnetized FM layers across NiMn indicates\na spin reorientation, possibly between an in-plane and an out-of-plane AFM spin structure\nat temperatures about halfway between the N\u0013 eel temperature and the onset of EB26. For\nthe present investigation we chose a stoichiometry close to Ni 40Mn60, the Ni to Mn ratio for\nwhich we have a detailed investigation of the N\u0013 eel-Temperature26. Both, the thickness and\nthe concentration, were chosen to have T Naround 360 K25{27. For the FM layer we selected\n3Ni with an out-of-plane and Co with an in-plane easy axis of magnetization.\nThe experiments are carried out at the femtoslicing facility at beamline UE56/1 ZPM of\nthe synchrotron radiation source BESSY II in Berlin. This facility provides ultra-fast soft\nX-ray pulses for time-resolved experiments28. The setup allows to measure the transmitted\nor re\rected intensity of the circularly polarized X-ray probe pulse following the pump laser\npulse. The system measures at 6 kHz repetition rate, alternating between X-ray probe\npulses with and without a leading laser pump pulse with variable pump{probe delay time.\nThis measurement scheme allows for quasi-simultaneous recording of the pumped and the\nunpumped signal. The X-ray and laser spot sizes are 140 \u0016m\u000240\u0016m and 1500 \u0016m\u0002200\n\u0016m, respectively. Both are co-propagating with an angle of 1 \u00002\u000ebetween them to separate\nthem again after re\rection at the sample. By switching the magnetization of the sample by\nan external magnetic \feld, we make use of the X-ray magnetic circular dichroism (XMCD)\nin re\rection29as an element-resolved probe of the magnetization. All measurements are\nperformed in saturation conditions under applied magnetic \felds of \u0006300 mT for Co/NiMn\nand\u0006400 mT for Ni/NiMn. The laser pulse has a temporal full width at half maximum\n(FWHM) of 60 fs and the X-ray pulse of 100 fs, with a resolving power of E/\u0001E = 500.\nTo \fnd the optimum measurement conditions, a range of static re\rectivity scans are\nperformed, identifying the angle with the highest magnetic contrast for a given photon\nenergy while still providing enough photon intensity for the slicing measurements. Energies\naround the L3edges of Ni and Co were checked and chosen such that the anticipated data\nacquisition time is minimized. For Co the optimum conditions are found at 7\u000eangle of\nincidence with respect to the surface and 775.6 eV photon energy, indicated in the re\rectivity\nand energy scans presented in Fig. 1. In a similar fashion, the optimum conditions for Ni\nare identi\fed at 6 :5\u000eangle of incidence and a photon energy of 845.0 eV.\nThe range of pump \ruences investigated is chosen such that the system completely de-\nmagnetizes at the highest \ruence and shows a considerable demagnetization at the lower\nend, in order to study the in\ruence of di\u000berent excitation strengths on the system. The\ncobalt system is thus investigated with incident \ruences between 10 and 50 mJ/cm2, cov-\nering a demagnetization range from 20 to 100%, as shown in Fig. 2. For Ni/NiMn we\nrecorded data for two di\u000berent incident \ruences of 15 and 46 mJ/cm2, resulting in 60 and\n90% relative demagnetization, respectively. From a layerwise absorption calculation the\nsurface-transmitted \ruence is estimated to be about 15-20 % of the incident \ruence, see\nappendix. To investigate the e\u000bect of the magnetic order of the antiferromagnetic layer on\n4(a)\n(b)\n6 8 10 12 1401234567\n+ M\n- M= 7°\nCu (001)NiMn (3.1 nm)Co (2.6 nm)Cu (3.1 nm) \n1.6 nm\n800 nm\n765 770 775 780 785\nIntensity (arb. units)\nInt. (arb. units)\n(deg.)Ephoton (eV)\nEphoton = 775.6 eVFIG. 1. (a) Schematic representation of the Co/NiMn sample. The measurements are performed in\nre\rectivity with an 800 nm pump and 775.6 eV ( \u00191:6 nm) probe pulse. (b) Optimum measurement\nconditions are found by taking spectra at di\u000berent angles and comparing \u0012\u00002\u0012scans at di\u000berent\nphoton energies. The most e\u000ecient conditions with respect to a minimized data acquisition time\nfor Co are found at \u0012= 7\u000e, highlighted by a vertical dashed line in the \fgure. The optimum photon\nenergy for Co is identi\fed as E= 775:6 eV, highlighted by the dashed vertical line shown in the\ninset of the \fgure, where a spectrum recorded at 7\u000eincidence is shown.\n5Time delay (ps)XMCD\n-0.25 0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.750.5\n0.4\n0.3\n0.2\n0.1\n0.010 mJ/cm2\n20 mJ/cm2\n30 mJ/cm2\n40 mJ/cm2\n50 mJ/cm2FIG. 2. Time delay traces of the Co demagnetization in Co/NiMn. Here the XMCD at the Co L3\nedge as a function of the delay at 390 K for di\u000berent laser \ruences is shown. With increased laser\n\ruence, it takes longer to reach the minimum magnetization, and also the remagnetization takes\nlonger. Complete demagnetization is reached with 50 mJ/cm2incident laser \ruence.\nthe demagnetization dynamics, two base temperatures of 80 and 390 K for Co/NiMn and\n80 and 360 K for Ni/NiMn were chosen, respectively. While the FM layer keeps its mag-\nnetization, the NiMn layer undergoes a phase transition from AFM at low temperatures to\nparamagnetic at the higher base temperature.\nIII. RESULTS\nFig. 2 shows delay traces of the time-dependent XMCD of the Co/NiMn sample, mea-\nsured at the Co L3edge, for 390 K sample temperature at di\u000berent incident pump \ruences.\nThe XMCD is de\fned as the asymmetry of the intensity of the re\rected beam with respect\nto the direction of the applied magnetic \feld. Upon excitation by the laser pulse, a de-\ncrease of the XMCD signal to a minimum in XMCD before 600 fs is observed in all cases.\n6With increasing \ruence the minimum shifts to longer times and the total change in XMCD\nincreases, following the expected behavior for ferromagnetic thin \flms after a higher heat\nload8. In order to compare the dynamics in Ni/NiMn and Co/NiMn, two measurements at\ndi\u000berent \ruences that show about the same relative demagnetization are presented in Fig.\n3. In the Co/NiMn sample [Fig. 3 (a)], the XMCD before time zero is virtually unchanged\nupon increase of the sample temperature due to the relatively high T Cof Co\u00181300 K at\n15 ML thickness30, decreasing only from 0.42 to 0.41. This is in contrast to the Ni/NiMn\nsample, where the XMCD decreases from 0.32 at 80 K to 0.18 at 360 K, shown in Fig. 3 (b)\nat negative delay times.\nWhile the static measurement does not reveal a di\u000berence between the two tempera-\ntures for Co/NiMn, the dynamic XMCD shows a di\u000berent time evolution. The decrease\nof magnetization of the Co layer at the lower base temperature, where the NiMn layer is\nantiferromagnetically ordered, is distinctly faster compared to the higher base temperature.\nTo evaluate this quantitatively, all delay traces recorded were \ftted using the sum of two\nexponentials describing a fast demagnetization and a subsequent slower initial remagneti-\nzation, respectively, from which the de- and remagnetization time constants are evaluated.\nThe sum of the exponentials is convoluted with a Gaussian function of FWHM = 120 fs to\naccount for the experimental resolution. The \ft function corresponds thus to\nM\nM0(t) =g(t)\n\u0010\nC\n+h\na(e\u0000t\u0000t0\n\u001cd\u00001)\u0000b(e\u0000t\u0000t0\n\u001cr\u00001)i\n\u0012(t\u0000t0)\u0011\n; (1)\nwith the initial XMCD C, the amplitudes for de- and remagnetization aandb, the de-\nand remagnetization time constants \u001cdand\u001cr, the Gaussian pro\fle g(t) of the instrumental\nresolution, and the Heaviside step function \u0012att0. The curves are \ftted with di\u000berent\namplitudes for de- and remagnetization, since the complete remagnetization dynamics may\nnot be adequately described by a single exponential. However, for the time window 2 ps\nused for the \ftting, the remagnetization, which is anyways not in the focus of the present\ninvestigation, can be well reproduced with a single exponential. As the relative timing of\npump and probe pulses shows slow drifts over the time period of this study, the curves\nplotted in this work are shifted by t0obtained from the \fts, in order to have a common zero\nfor all graphs.\nA \ruence-dependent comparison of the demagnetization times and amplitudes for dif-\nferent temperatures is presented in Fig. 4. For both base temperatures and samples, the\n7Time delay (ps)XMCD XMCD0.6\n0.4\n0.2\n(a)\n(b)Co/NiMn\nNi/NiMn80 K\n390 K\n0.4\n0.3\n0.2\n0.1\n0.080 K\n360 K\n-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 FIG. 3. Selected time delay traces for the demagnetization of Co and Ni in Co/NiMn and Ni/NiMn,\nrespectively. Red circles indicate the high and blue squares the low base temperature, as indicated\nin the legends; the \flled symbols show data for the Co/NiMn sample and open symbols for the\nNi/NiMn sample. Panel (a) shows delay traces of the demagnetization of Co at 30 mJ/cm2laser\n\ruence at 80 and 390 K, panel (b) the demagnetization of Ni/NiMn after laser excitation with 15\nmJ/cm2laser \ruence.\nquench of the magnetization, depicted in Fig. 4 (a) as change of XMCD \u0001 XMCD in the same\nabsolute values as on the vertical axes in Figs. 2 and 3, increases with \ruence. Except for\nthe highest \ruence, this is true also for the demagnetization time \u001cd. For the higher sample\ntemperature, demagnetization times are longer and the relative demagnetization stronger,\nsimilar to the observations in Ref.31. In absolute values, i.e., in XMCD asymmetry \u0001 XMCD ,\nthe cold Ni/NiMn sample demagnetizes more than the warm one, but when considering the\n8∆XMCD (fs)τd\n10 20 30 40 50200\n180\n160\n140\n120\n100\n80\n60\n400.40\n0.35\n0.30\n0.25\n0.20\n0.15\n0.10\nFluence (mJ/cm2)(a)\n(b)\nCo 80 K\nCo 390 K\nNi 80 K\nNi 360 K\nFIG. 4. (a) Demagnetization amplitude in XMCD asymmetry after laser excitation for Co (\flled\nsymbols) and Ni (open symbols), at 80 K and 360 K for Ni and 390 K for Co, respectively, as\ndenoted in the legend in panel (b). The amplitude of the demagnetization follows a linear trend at\nlower \ruences and is slightly higher for elevated sample temperatures for cobalt, considering the\nsame \ruence. (b) Magnetization decay time \u001cdversus \ruence. Generally, a higher base temperature\nleads to longer demagnetization times with the largest di\u000berence for Co/NiMn around 30 to 40\nmJ/cm2laser \ruence.\nrelative amount of magnetization quenched at the lower temperature it is actually less than\nat the higher temperature. When the Co/NiMn system fully demagnetizes, which is the\ncase for 50 mJ/cm2at 390 K, the demagnetization is again faster.\nTo obtain more information on the role of the antiferromagnetic order in the NiMn layer,\nwe compare the magnetization decay times for Co and Ni as a function of the reduction of\nthe XMCD signal in Fig. 5. For Co/NiMn, the \ruence range covers similar demagnetization\n9(fs)τd\n∆XMCD200\n180\n160\n140\n120\n100\n80\n60\n40\n0.35 0.30 0.25 0.20 0.15 0.10Co 80 K\nCo 390 K\nNi 80 K\nNi 360 K\nFIG. 5. Demagnetization times \u001cdvs. demagnetization amplitude for Co/NiMn (\flled symbols)\nand Ni/NiMn (open symbols) at 80 K (blue) and 360 K for Ni and 390 K for Co, respectively (red).\nThe demagnetization time, considering a similar quench of XMCD asymmetry, shows a di\u000berence\nfor the two temperatures measured. In Co/NiMn, for both base temperatures an XMCD of 0.42 for\n80 K and 0.41 for 390 K was measured before time zero, while for Ni/NiMn an XMCD asymmetry\nof 0.32 for 80 K and 0.18 for 360 K was observed at negative delay times.\nat both temperatures, while for Ni/NiMn, where the static di\u000berence in magnetization is\nalready quite sizeable, this is not the case. We observe a di\u000berence of \u001cdof up to 50 fs for\nCo/NiMn for the two temperatures around a demagnetization corresponding to a change\nof the XMCD of 0.3. The di\u000berence, albeit small, is still greater than the experimental\nerror and the uncertainty from \ftting. The data for Ni/NiMn suggest that the di\u000berence\nin demagnetization times for demagnetization amplitudes around 0.15{0.20 could be even\ngreater than that.\n10IV. DISCUSSION\nThe comparison of the demagnetization time constants at di\u000berent temperatures for equal\ndemagnetization amplitudes (Fig. 5) highlights the role of the AFM order in the NiMn layer\nfor the speed of demagnetization of the FM layer. The results show that the demagnetization\nrate of the FM layer is higher if the AFM layer exhibits magnetic order. It makes the\ninclusion of new dissipation channels introduced by the adjacent AFM layer necessary for the\ndescription of the demagnetization dynamics. We suggest in the following that superdi\u000busive\nspin currents15{18are responsible for the faster demagnetization of the FM layer at the lower\ntemperature, i.e., in the presence of antiferromagnetic order in the NiMn layer.\nTransport by superdi\u000busive currents plays an important role for the ultrafast demagneti-\nzation as shown, for example, in Refs.16{18, where a contribution of transport to an adjacent\nlayer to the ultrafast demagnetization has been reported. Before we discuss superdi\u000busive\nspin currents in the sample, we take a short look at the simulated di\u000berential absorbed laser\n\ruence across the Co/NiMn bilayer, presented in the Appendix, Fig. 6. The energy absorbed\nin the Co layer is about the same as the one absorbed in the NiMn layer, with a di\u000berence\nof only about 1{2%. Hot electrons will thus be excited both in the Co and in the NiMn\nlayer. Excitations in the FM Co layer give rise to a superdi\u000busive spin current that enters\nthe NiMn layer as well as the cap layer. Similarly, the excitation of the NiMn layer will lead\nto superdi\u000busive hot electrons that will be transported through the interface into the FM\nlayer and the Cu substrate. We will discuss in the following how these superdi\u000busive spin\ncurrents contribute to the observed di\u000berence in FM-layer demagnetization times below and\nabove the AFM ordering temperature.\nThe decisive aspect for the demagnetization of the FM layer is the dissipation of angular\nmomentum by the exchange of spin currents with the NiMn layer. Although spin mixing is\nenhanced after pumping, the superdi\u000busive spin current carried by hot electrons from the\nFM layer consists mainly of spin-majority species15. BelowTN, we assume the magnetization\nin the FM layer to be collinear with the magnetization axis of the sublattices of the ordered\nAFM NiMn layer. This con\fguration results in an injection of the spin-polarized electrons\ninto empty states of the AFM of the same spin character. In the high-temperature param-\nagnetic state, though, the magnetic moments in the NiMn layer are disordered, resulting in\na spin-character mismatch due to the random relative orientation of the spin moments in\nthe NiMn layer with respect to the FM-aligned spins in Co. Under such conditions, the spin\n11penetration depth becomes strongly reduced to about one nanometer32, leading to re\rection\nof carriers at the interface. Conversely, spin-polarized electron penetration into magnetic\nsystems with collinear spin orientation is much larger33. In the AFM-ordered state with a\ncollinear spin structure, the superdi\u000busive majority-type spin current from the FM layer can\nthus propagate much farther into the NiMn layer, making it a more e\u000ecient sink for angular\nmomentum. This leads to shorter demagnetization times for the FM layer adjacent to the\nordered AFM layer, as is observed experimentally.\nIn turn, the superdi\u000busive current of hot electrons originating in the NiMn layer will be\ntransported through the interface into the FM layer. Likewise, for collinear AFM spin order\nin the NiMn layer, interface re\rection is reduced compared to the case of a magnetically\ndisordered NiMn layer. Despite the absence of a macroscopic magnetization in an AFM,\na spin current representing the minority species of the FM layer will be preferentially en-\ntering the FM layer, as there are more unoccupied minority-spin states available, thus also\naccelerating its demagnetization.\nWe now argue that the change of temperature alone, without considering the ordering\ntransition in the NiMn layer, cannot explain our experimental observation. The e\u000bect of\ntemperature has not yet been discussed in detail in the model of superdi\u000busive transport\n(see15and34), where the FM layer is considered fully magnetic. In general, spin lifetimes and\nspin-dependent velocities of the excited electrons change with temperature. This could play\na role in the Ni/NiMn sample, where the asymmetry in the spin lifetimes and spin velocities\nwill be reduced at 360 K, in the same way as the static XMCD. For Co, though, a change\nof this asymmetry is negligible since the change of the experimentally measured XMCD is\nmarginal between the two temperatures.\nIn addition, the excited electron lifetime decreases at higher temperatures due to enhanced\nscattering probabilities. Furthermore, slightly di\u000berent states will be addressed at di\u000berent\ntemperatures, depending on the slope of the band structure. This could in\ruence the electron\nvelocity in either direction, but this is also considered a negligible e\u000bect here as the change\nin temperature is only about 300 K. Following this line of argumentation, we do not expect\na signi\fcant change in the demagnetization time for equal demagnetization as a result of\nthe di\u000berent base temperature in the FM.\nNext we exclude local mechanisms for angular momentum dissipation as a possible ex-\nplanation. Local demagnetization can be caused by Elliott-Yafet electron{phonon spin-\rip\nscattering8or ultrafast magnon generation14,35{37. In the work of Schellekens et al.38, lo-\n12cal demagnetization due to Elliott-Yafet spin-\rip scattering has been proposed as the main\ndriving mechanism that could be orders of magnitude stronger compared to spin transport.\nHowever, the relative contributions of the microscopic processes strongly depend on the sys-\ntem and properties under investigation14,39. While electronic-structure-based calculations of\nthe Elliott-Yafet electron{phonon spin-\rip scattering predict that this can only give a small\ncontribution to the laser-induced demagnetization40,41, others suggested that a feedback\nmechanism of the magnetization change and the spin-dependent chemical potential on the\nband structure would lead to an additional reduction of the exchange splitting42,43. However,\nsuch a collapse of the exchange splitting has not been found in femtosecond spin-polarized\nphotoemission experiments36and also not in extreme ultraviolet transversal magneto-optical\nKerr e\u000bect (TMOKE) measurements14on Co. In any case, using the expressions given in\nRef.40one can calculate that an increase in temperature from 80 to 390 K leads to an in-\ncrease in the Elliott-Yafet spin-\rip scattering by about a factor of 2.5 due to the increased\nphonon population. This corresponds to a faster electron{phonon demagnetization of Co at\nhigher temperatures, which is, however, opposite to the here-measured trend, which shows\naslower demagnetization of Co in Co/NiMn at 390 K.\nMagnetization dissipation by transversal spin excitations, i.e., ultrafast magnon generation14,36,37,\nis another mechanism we exclude, based on the timescale of our observations. Analogous to\nthe case of ultrafast magnon generation followed by spin pumping into an adjacent layer44,45,\nmagnons excited in the FM layer by ultrafast carriers could couple to magnon modes in the\nordered AFM layer, a coupling that would be absent in the disordered case. This could also\nlead to a faster demagnetization of the FM layer when in contact with the ordered AFM\nlayer. However, the timescale of the initial energy transfer process from hot electrons to hot\nmagnons in an FM has not been de\fnitely established. E\u000bects of hot magnon excitation\nhave been detected in laser-excited Co at already 100 fs36. The timescale to couple the hot\nFM magnons to magnon modes in the AFM is neither precisely known. Magnon transport,\nassuming even a quite high magnon group velocity of \u0018104m/s46, would only result in an\nangular momentum transfer over 1 nm in 100 fs, which is not enough to explain the faster\ndemagnetization timescale seen in Fig. 5. Thus, although the angular momentum transfer\nprocess, starting with hot magnons in the FM layer and their subsequent di\u000busion into the\nadjacent layer, is possible, it is expected to occur on a longer timescale than transfer by\nsuperdi\u000busive spin currents.\nWe \fnally note that, although our data from the Ni/NiMn sample do not allow the\n13comparison of demagnetization times for the two temperatures at equal demagnetization\namplitudes, we expect the same e\u000bect as in the Co/NiMn sample. This is motivated by the\nNi data presented in Fig. 5, where one pair of data points of almost equal magnetization\nquench shows a much faster rate at the lower temperature.\nV. CONCLUSIONS AND OUTLOOK\nIn conclusion, we infer from our experimental \fndings superdi\u000busive spin transport to\nplay a signi\fcant role for the demagnetization of an FM/AFM system. A general temper-\nature dependence of demagnetization as well as the contribution of Elliott-Yafet processes\ncannot account for the observed temperature dependence of the demagnetization. More-\nover, a magnon-related contribution to the demagnetization is expected to occur on a longer\ntimescale. Therefore, we attribute the accelerated demagnetization at low temperatures in\nCo/NiMn and Ni/NiMn to the presence of AFM order in the NiMn layer.\nSuperdi\u000busive spin currents can be exchanged more easily between the FM layer and the\nadjacent AFM-ordered NiMn layer, while the disordered NiMn layer acts as a barrier for\nthe penetration of superdi\u000busive currents. The AFM-ordered NiMn layer thus represents\non the one hand a sink for majority spin current from the FM layer, while on the other\nhand it injects minority spin current into the FM layer. Both accelerates the FM-layer\ndemagnetization and explains the experimental observations.\nTo accelerate the ultrafast demagnetization by AFM spin order in an adjacent layer\nmay become important for the ultrafast optical manipulation of magnetic order in magnetic\nrecording or spintronic devices. Rotating the spin axis of the AFM layer between perpen-\ndicular and parallel to the FM-layer magnetization, for example by staggered spin-orbit\ntorques47, could be a means of tuning the demagnetization speed of an FM layer. It might\n\fnally be interesting to see, possibly by time-resolved magnetic linear dichroism, how in a\nreverse experiment the presence of FM spin order in an adjacent layer in\ruences the optical\nquench of the magnetic order in an AFM layer.\nACKNOWLEDGMENTS\nThis work was supported by the Deutsche Forschungsgemeinschaft via the CRC/TRR\n227 \\Ultrafast Spin Dynamics\", projects A03 and A07, and by the Swedish Research Council\n14(VR). We further acknowledge support from the K. and A. Wallenberg Foundation (grant\nNo. 2015.0060) and the Swedish National Infrastructure for Computing (SNIC). We thank\nthe Helmholtz-Zentrum Berlin for the allocation of synchrotron radiation beamtime.\nAppendix A: Di\u000berential Absorption\nAbsorption (% nm-1)\nDistance from surface (nm)3.0\n2.5\n2.0\n1.5\n1.0\n0.5\n0.0\n0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5\nCu\nCapCu\nSubstrateCo \nNiMn\n7.0 %\n6.9 %7.1 %\n6.7 %total absorbed\n power\n16.6 %\n16.1 %nNiMn = nNi\nnNiMn = nMn\nFIG. 6. Di\u000berential absorption over sample depth for the Co/NiMn sample. The optical constants\nfor NiMn are estimated to lie between the ones for bulk Ni and for bulk Mn. To estimate the\nabsorbed energy in each layer, two absorption pro\fles are shown, one with the optical constants of\nNi for the NiMn alloy (red) and one with the optical constants of Mn (blue). The numbers indicate\nthe fraction of absorbed pump-pulse \ruence in each layer.\nWe simulate the layerwise absorption of the pump pulse in the Co/NiMn sample using the\nmatrix formalism described in Ref.48. The optical constants of NiMn alloy are approximated\nby calculating the boundary scenarios of a pure Ni or Mn \flm. To estimate the absorption,\nboth cases are presented in Fig. 6, where the NiMn alloy absorption is calculated once\nwith the constants of Ni as nNi= 2:2180 +i4:8925 and once with the ones of Mn, nMn=\n2:7880 +i3:9982, both taken from Ref.49. The vertical layer distance for NiMn has been\n15measured at the present concentration as 1 :88\u0017A/ML25, for Co we use 1.74 \u0017A/ML50and for\nCu 1:81\u0017A/ML. In both cases the absorption in the Co layer is slightly higher, but each layer\nabsorbs roughly the same magnitude, \u00197%, of the incident \ruence.\n1C. H. Back, D. Weller, J. Heidmann, D. Mauri, D. Guarisco, E. L. Garwin, and H. C. Siegmann,\nMagnetization Reversal in Ultrashort Magnetic Field Pulses , Phys. Rev. Lett. 81, 3251 (1998).\n2H. W. Schumacher, C. Chappert, R. C. Sousa, P. P. Freitas, and J. Miltat, Quasiballistic\nMagnetization Reversal , Phys. Rev. Lett. 90, 017204 (2003).\n3I. Tudosa, C. Stamm, A. B. Kashuba, F. King, H. C. Siegmann, J. St ohr, G. Ju, B. Lu, and\nD. Weller, The ultimate speed of magnetic switching in granular recording media , Nature 428,\n831 (2004).\n4E. Beaurepaire, J.-C. Merle, A. Daunois, and J.-Y. Bigot, Ultrafast spin dynamics in ferromag-\nnetic nickel , Phys. Rev. Lett. 76, 4250 (1996).\n5B. Koopmans, J. J. M. Ruigrok, F. D. Longa, and W. J. M. de Jonge, Unifying Ultrafast\nMagnetization Dynamics , Phys. Rev. Lett. 95, 267207 (2005).\n6M. Cinchetti, M. S\u0013 anchez Albaneda, D. Ho\u000bmann, T. Roth, J.-P. W ustenberg, M. Krauss,\nO. Andreyev, H. C. Schneider, M. Bauer, and M. Aeschlimann, Spin-\rip processes and ultrafast\nmagnetization dynamics in Co: Unifying the microscopic and macroscopic view of femtosecond\nmagnetism , Phys. Rev. Lett. 97, 177201 (2006).\n7J.-Y. Bigot, M. Vomir, and E. Beaurepaire, Coherent ultrafast magnetism induced by femtosec-\nond laser pulses , Nat. Phys. 5, 515 (2009).\n8B. Koopmans, G. Malinowski, F. Dalla Longa, D. Steiauf, M. F ahnle, T. Roth, M. Cinchetti, and\nM. Aeschlimann, Explaining the paradoxical diversity of ultrafast laser-induced demagnetization ,\nNat. Mater. 9, 259 (2010).\n9A. Kirilyuk, A. V. Kimel, and T. Rasing, Ultrafast optical manipulation of magnetic order , Rev.\nMod. Phys. 82, 2731 (2010).\n10C.-H. Lambert, S. Mangin, B. S. D. C. S. Varaprasad, Y. K. Takahashi, M. Hehn, M. Cinchetti,\nG. Malinowski, K. Hono, Y. Fainman, M. Aeschlimann, and E. E. Fullerton, All-optical control\nof ferromagnetic thin \flms and nanostructures , Science 345, 1337 (2014).\n11J. Walowski and M. M unzenberg, Perspective: Ultrafast magnetism and THz spintronics , J.\nAppl. Phys. 120, 140901 (2016).\n1612C. Stamm, T. Kachel, N. Pontius, R. Mitzner, T. Quast, K. Holldack, S. Khan, C. Lupulescu,\nE. F. Aziz, M. Wietstruk, H. A. D urr, and W. Eberhardt, Femtosecond modi\fcation of electron\nlocalization and transfer of angular momentum in nickel , Nat. Mater. 6, 740 (2007).\n13M. Krau\u0019, T. Roth, S. Alebrand, D. Steil, M. Cinchetti, M. Aeschlimann, and H. C. Schnei-\nder, Ultrafast demagnetization of ferromagnetic transition metals: The role of the Coulomb\ninteraction , Phys. Rev. B 80, 180407 (2009).\n14E. Turgut, D. Zusin, D. Legut, K. Carva, R. Knut, J. M. Shaw, C. Chen, Z. Tao, H. T. Nembach,\nT. J. Silva, S. Mathias, M. Aeschlimann, P. M. Oppeneer, H. C. Kapteyn, M. M. Murnane, and\nP. Grychtol, Stoner versus Heisenberg: Ultrafast exchange reduction and magnon generation\nduring laser-induced demagnetization , Phys. Rev. B 94, 220408 (2016).\n15M. Battiato, K. Carva, and P. M. Oppeneer, Superdi\u000busive spin transport as a mechanism of\nultrafast demagnetization , Phys. Rev. Lett. 105, 027203 (2010).\n16A. Eschenlohr, M. Battiato, P. Maldonado, N. Pontius, T. Kachel, K. Holldack, R. Mitzner,\nA. F ohlisch, P. M. Oppeneer, and C. Stamm, Ultrafast spin transport as key to femtosecond\ndemagnetization , Nat. Mater. 12, 332 (2013).\n17D. Rudolf, C. La-o Vorakiat, M. Battiato, R. Adam, J. M. Shaw, E. Turgut, P. Maldonado,\nS. Mathias, P. Grychtol, H. T. Nembach, T. J. Silva, M. Aeschlimann, H. C. Kapteyn, M. M.\nMurnane, C. M. Schneider, and P. M. Oppeneer, Ultrafast magnetization enhancement in metal-\nlic multilayers driven by superdi\u000busive spin current , Nat. Commun. 3, 1037 (2012).\n18E. Turgut, C. La-o vorakiat, J. M. Shaw, P. Grychtol, H. T. Nembach, D. Rudolf, R. Adam,\nM. Aeschlimann, C. M. Schneider, T. J. Silva, M. M. Murnane, H. C. Kapteyn, and S. Mathias,\nControlling the Competition between Optically Induced Ultrafast Spin-Flip Scattering and Spin\nTransport in Magnetic Multilayers , Phys. Rev. Lett. 110, 197201 (2013).\n19W. H. Meiklejohn and C. P. Bean, New Magnetic Anisotropy , Phys. Rev. 102, 1413 (1956).\n20J. Nogu\u0013 es and I. K. Schuller, Exchange bias , J. Magn. Magn. Mater. 192, 203 (1999).\n21G. Ju, A. V. Nurmikko, R. F. C. Farrow, R. F. Marks, M. J. Carey, and B. A. Gurney, Ultrafast\noptical modulation of an exchange biased ferromagnetic/antiferromagnetic bilayer , Phys. Rev.\nB58, 11857 (1998).\n22G. Ju, A. V. Nurmikko, R. F. C. Farrow, R. F. Marks, M. J. Carey, and B. A. Gurney, Ultrafast\nTime Resolved Photoinduced Magnetization Rotation in a Ferromagnetic/Antiferromagnetic\nExchange Coupled System , Phys. Rev. Lett. 82, 3705 (1999).\n23G. Ju, L. Chen, A. V. Nurmikko, R. F. C. Farrow, R. F. Marks, M. J. Carey, and\n17B. A. Gurney, Coherent magnetization rotation induced by optical modulation in ferromag-\nnetic/antiferromagnetic exchange-coupled bilayers , Phys. Rev. B 62, 1171 (2000).\n24F. Dalla Longa, J. T. Kohlhepp, W. J. M. de Jonge, and B. Koopmans, Resolving the gen-\nuine laser-induced ultrafast dynamics of exchange interaction in ferromagnet/antiferromagnet\nbilayers , Phys. Rev. B 81, 094435 (2010).\n25C. Tieg, W. Kuch, S. G. Wang, and J. Kirschner, Growth, structure, and magnetism of single-\ncrystalline NixMn 100\u0000x\flms andNiMnCo bilayers on Cu(001) , Phys. Rev. B 74, 094420\n(2006).\n26T. Hagelschuer, Y. A. Shokr, and W. Kuch, Spin-state transition in antiferromagnetic\nNi0:4Mn 0:6\flms in Ni/NiMn/Ni trilayers on Cu(001) , Phys. Rev. B 93, 054428 (2016).\n27M. Reinhardt, J. Seifert, M. Busch, and H. Winter, Magnetic interface coupling between ultrathin\nCo andNixMn 100\u0000x\flms on Cu(001) , Phys. Rev. B 81, 134433 (2010).\n28K. Holldack, J. Bahrdt, A. Balzer, U. Bovensiepen, M. Brzhezinskaya, A. Erko, A. Eschenlohr,\nR. Follath, A. Firsov, W. Frentrup, L. Le Guyader, T. Kachel, P. Kuske, R. Mitzner, R. M uller,\nN. Pontius, T. Quast, I. Radu, J. S. Schmidt, C. Sch ussler-Langeheine, M. Sperling, C. Stamm,\nC. Trabant, and A. F ohlisch, FemtoSpeX: a versatile optical pump-soft X-ray probe facility with\n100 fs X-ray pulses of variable polarization , J. Synchr. Rad. 21, 1090 (2014).\n29H.-C. Mertins, D. Abramsohn, A. Gaupp, F. Sch afers, W. Gudat, O. Zaharko, H. Grimmer,\nand P. M. Oppeneer, Resonant magnetic re\rection coe\u000ecients at the Fe 2pedge obtained with\nlinearly and circularly polarized soft X-rays , Phys. Rev. B 66, 184404 (2002).\n30C. M. Schneider, P. Bressler, P. Schuster, J. Kirschner, J. J. de Miguel, and R. Miranda, Curie\ntemperature of ultrathin \flms of fcc-cobalt epitaxially grown on atomically \rat Cu(100) surfaces ,\nPhys. Rev. Lett. 64, 1059 (1990).\n31T. Roth, A. J. Schellekens, S. Alebrand, O. Schmitt, D. Steil, B. Koopmans, M. Cinchetti, and\nM. Aeschlimann, Temperature Dependence of Laser-Induced Demagnetization in Ni: A Key for\nIdentifying the Underlying Mechanism , Phys. Rev. X 2, 021006 (2012).\n32A. Ghosh, S. Au\u000bret, U. Ebels, and W. E. Bailey, Penetration Depth of Transverse Spin Current\nin Ultrathin Ferromagnets , Phys. Rev. Lett. 109, 127202 (2012).\n33A. Alekhin, I. Razdolski, N. Ilin, J. P. Meyburg, D. Diesing, V. Roddatis, I. Rungger, M. Stamen-\nova, S. Sanvito, U. Bovensiepen, and A. Melnikov, Femtosecond Spin Current Pulses Generated\nby the Nonthermal Spin-Dependent Seebeck E\u000bect and Interacting with Ferromagnets in Spin\nValves , Phys. Rev. Lett. 119, 017202 (2017).\n1834M. Battiato, K. Carva, and P. M. Oppeneer, Theory of laser-induced ultrafast superdi\u000busive\nspin transport in layered heterostructures , Phys. Rev. B 86, 024404 (2012).\n35E. Carpene, E. Mancini, C. Dallera, M. Brenna, E. Puppin, and S. De Silvestri, Dynamics of\nelectron-magnon interaction and ultrafast demagnetization in thin iron \flms , Phys. Rev. B 78,\n174422 (2008).\n36S. Eich, M. Pl otzing, M. Rollinger, S. Emmerich, R. Adam, C. Chen, H. C. Kapteyn, M. M.\nMurnane, L. Plucinski, D. Steil, B. Stadtm uller, M. Cinchetti, M. Aeschlimann, C. M. Schneider,\nand S. Mathias, Band structure evolution during the ultrafast ferromagnetic-paramagnetic phase\ntransition in cobalt , Sci. Adv. 3, e1602094 (2017).\n37R. Gort, K. B uhlmann, S. D aster, G. Salvatella, N. Hartmann, Y. Zemp, S. Holenstein,\nC. Stieger, A. Fognini, T. U. Michlmayr, T. B ahler, A. Vaterlaus, and Y. Acremann, Early\nStages of Ultrafast Spin Dynamics in a 3d Ferromagnet , Phys. Rev. Lett. 121, 087206 (2018).\n38A. J. Schellekens, W. Verhoeven, T. N. Vader, and B. Koopmans, Investigating the contribution\nof superdi\u000busive transport to ultrafast demagnetization of ferromagnetic thin \flms , Appl. Phys.\nLett. 102, 252408 (2013).\n39J. Wieczorek, A. Eschenlohr, B. Weidtmann, M. R osner, N. Bergeard, A. Tarasevitch, T. O.\nWehling, and U. Bovensiepen, Separation of ultrafast spin currents and spin-\rip scattering in\nCo/Cu(001) driven by femtosecond laser excitation employing the complex magneto-optical Kerr\ne\u000bect , Phys. Rev. B 92, 174410 (2015).\n40K. Carva, M. Battiato, and P. M. Oppeneer, Ab Initio Investigation of the Elliott-Yafet Electron-\nPhonon Mechanism in Laser-Induced Ultrafast Demagnetization , Phys. Rev. Lett. 107, 207201\n(2011).\n41S. Essert and H. C. Schneider, Electron-phonon scattering dynamics in ferromagnetic metals\nand their in\ruence on ultrafast demagnetization processes , Phys. Rev. B 84, 224405 (2011).\n42A. J. Schellekens and B. Koopmans, Comparing ultrafast demagnetization rates between com-\npeting models for \fnite temperature magnetism , Phys. Rev. Lett. 110, 217204 (2013).\n43B. Y. Mueller, A. Baral, S. Vollmar, M. Cinchetti, M. Aeschlimann, H. C. Schneider, and\nB. Rethfeld, Feedback E\u000bect during Ultrafast Demagnetization Dynamics in Ferromagnets , Phys.\nRev. Lett. 111, 167204 (2013).\n44G.-M. Choi, B.-C. Min, K.-J. Lee, and D. G. Cahill, Spin current generated by thermally driven\nultrafast demagnetization , Nat. Commun. 5, 4334 (2014).\n45G.-M. Choi and D. G. Cahill, Kerr rotation in Cu, Ag, and Au driven by spin accumulation\n19and spin-orbit coupling , Phys. Rev. B 90, 214432 (2014).\n46X. Wu, Z. Liu, and T. Luo, Magnon and phonon dispersion, lifetime, and thermal conductivity\nof iron from spin-lattice dynamics simulations , J. Appl. Phys. 123, 085109 (2018).\n47P. Wadley, B. Howells, J. \u0014Zelezn\u0013 y, C. Andrews, V. Hills, R. P. Campion, V. Nov\u0013 ak,\nK. Olejn\u0013 \u0010k, F. Maccherozzi, S. S. Dhesi, S. Y. Martin, T. Wagner, J. Wunderlich, F. Freimuth,\nY. Mokrousov, J. Kune\u0014 s, J. S. Chauhan, M. J. Grzybowski, A. W. Rushforth, K. W. Edmonds,\nB. L. Gallagher, and T. Jungwirth, Electrical switching of an antiferromagnet , Science 351, 587\n(2016).\n48K. Ohta and H. Ishida, Matrix formalism for calculation of electric \feld intensity of light in\nstrati\fed multilayered \flms , Appl. Opt. 29, 1952 (1990).\n49P. B. Johnson and R. W. Christy, Optical constants of transition metals: Ti, V, Cr, Mn, Fe,\nCo, Ni, and Pd , Phys. Rev. B 9, 5056 (1974).\n50J. R. Cerd\u0013 a, P. L. de Andres, A. Cebollada, R. Miranda, E. Navas, P. Schuster, C. M. Schnei-\nder, and J. Kirschner, Epitaxial growth of cobalt \flms on Cu(100): a crystallographic LEED\ndetermination , J. Phys.: Condens. Matt. 5, 2055 (1993).\n20" }, { "title": "1009.5798v2.Acoustically_driven_ferromagnetic_resonance.pdf", "content": "arXiv:1009.5798v2 [cond-mat.mtrl-sci] 6 Apr 2011Elastically driven ferromagnetic resonance in nickel thin films\nM. Weiler,1L. Dreher,2C. Heeg,1H. Huebl,1R.\nGross,1M.S. Brandt,2and S.T.B. Goennenwein1,∗\n1Walther-Meißner-Institut, Bayerische Akademie\nder Wissenschaften, 85748 Garching, Germany\n2Walter Schottky Institut, Technische Universit¨ at M¨ unch en, 85748 Garching, Germany\nAbstract\nSurface acoustic waves (SAW) in the GHz frequency range are e xploited for the all-elastic excita-\ntion and detection of ferromagnetic resonance (FMR) in a fer romagnetic/ferroelectric (Ni/LiNbO 3)\nhybrid device. We measure the SAW magneto-transmission at r oom temperature as a function of\nfrequency, external magnetic field magnitude, and orientat ion. Our data are well described by a\nmodified Landau-Lifshitz-Gilbert approach, in which a virt ual, strain-induced tickle field drives the\nmagnetization precession. This causes a distinct magnetic field orientation dependence of elastically\ndriven FMR that we observe in both model and experiment.\n1Inverse magnetostriction, or magnetoelasticity, enables the con trol of the magnetization in\nferromagnetic materials via elastic stress [1–6]. This spin-mechanica l interaction prevails at\nradio frequencies (RF), so that magnonic and phononic degrees of freedom become coupled, as\ndiscussed theoretically [7–10]. We here apply this concept to a surfa ce acoustic wave (SAW)\ntraversing a ferromagnetic thin film. Due to magnetoelastic coupling [11], the elastic defor-\nmation periodic in time and space results in a change of the magnetic an isotropy, which in\nturn exerts a torque on the magnetization. Since typical SAW freq uencies range from a few\nMHz to several GHz, SAW-based RF spin-mechanics should enable th e study of magnetization\ndynamics such as ferromagnetic resonance, driven only via RF elast ic deformation, not via ex-\nternally applied RF magnetic fields. The interaction of SAWs and ferro magnetic thin films has\nbeen studied experimentally by several groups [12–15]. Based on SA W magneto-transmission\nmeasurements with the static external magnetic field applied either perpendicular or parallel\nto the SAW propagation direction, these authors suggested that a magnetoelastic interaction\nmost probably was the dominant interaction mechanism. However, c onclusive evidence for the\noccurrence of an elastically driven, acoustic FMR has proven elusive , and important aspects of\nthe interaction mechanism still await explanation, as stated by Wiege rt as recently as 2002 [16].\nOur experimental study of SAW-based RF spin mechanics as a funct ion of magnetic field mag-\nnitude and orientation provides clear evidence for elastically excited , acoustic ferromagnetic\nresonance. Our findings thus extend the application and understa nding of magnetoelastic in-\nteraction phenomena in the RF regime.\nWe study the hybrid device depicted schematically in Fig. 1(a). It con sists of a 50nm\nthick polycrystalline ferromagnetic nickel film deposited onto the ce ntral part of a 5 ×6mm2\ny-cut lithium niobate substrate. 70nm thick Aluminum interdigital tra nsducers (IDTs) [17]\nwith a finger width of 5 µm are employed to launch and detect a SAW. The complex forward\ntransmission S21oftheapproximately850 µmlongdelaylineisdeterminedusingvectornetwork\nanalysis with a time domain window to cancel the contributions of elect romagnetic crosstalk\nand triple transit interference from the signal. For the description of the experimental and\ntheoretical results we employ the cartesian coordinate system sh own in Fig. 1(a), where αis\nthe angle between an external magnetic field Happlied in the Ni film plane and the SAW\npropagation direction kSAW∝ba∇dblx. We first discuss the influence of the SAW on the magnetization\nvectorMbasedonmagnetoelasticcoupling. Inthestaticlimit, themagneticfr eeenergydensity\n2FIG. 1. (color online) Principles of elastically driven fer romagnetic resonance. (a) Illustration of\nexperimental setup and coordinate system. The closeup to th e right shows the strain εin the Nickel\nthin film. (b) The color coded caps show calculations of the ma gnetic free energy density F. The\nblack arrows depict the gradient of Fwith respect to the magnetization direction m. A finite gradient\nat the equilibrium orientation of M(red arrow) and thus FMR is found only for α=−45◦andε∝ne}ationslash= 0.\nF0of the Ni film normalized to the saturation magnetization Msis given by [18]\nF0(m) =−µ0H·m+Bdm2\nz+Bu(u·m)2+const. , (1)\nwherem=M/Mswith components {mx,my,mz}.Bd=µ0Ms/2 is the shape anisotropy and\nBuis a uniaxial in-plane anisotropy field along u={ux,uy,0}. In equilibrium, the magnetiza-\ntion is oriented along a minimum of F0. Due to the shape anisotropy, and as we only consider\nHin the film plane here, the equilibrium morientation is in the film plane, at an angle θ0\nbetween xandm.\nThe SAW generates an RF strain εin the ferromagnetic thin film plane. Due to magne-\ntoelasticity [18], this strain results in an RF contribution FRFto the magnetic free energy\ndensity\nFRF(m) =B1ε(x,t)m2\nx+const. , (2)\nwhereB1is the magnetoelastic coupling constant and shear strains have bee n neglected for\nsimplicity. The effective field Heffacting on mis given by −∇mF=−∇m/parenleftbig\nF0+FRF/parenrightbig\n[19, 20],\n3FIG. 2. (color online) Experimental observation of elastic ally driven FMR in SAW transmission. (a)\n|S21|as a function of frequency at H= 0. The inset shows |S21|around the 9thharmonic, recorded\nwith the external magnetic field applied at α= 30◦, forµ0|H|= 100mT (solid line) and µ0|H|= 7mT\n(dotted line). (b) |S21|as a function of the external magnetic field applied at α= 30◦(solid lines:\nupsweep, dashed lines: downsweep). With increasing freque ncy, the damping of the SAW increases\nand the field-symmetric absorption dips shift to higher magn etic fields, as expected for FMR.\nevaluated at the equilibrium orientation θ0.\nThe qualitative effect of the SAW on the magnetization is illustrated in F ig. 1(b), where we\nexemplarilyconsider anin-planemagneticallyisotropicnickel film( Bu= 0). Thetotalmagnetic\nfree energy density Fis depicted by the color code in the caps close to the magnetization\nequilibrium position θ0=αforα∈ {−90◦,−45◦,0◦}andε <0,ε= 0 and ε >0. The ”tickle\nfield”, i.e., the components of Heffperpendicular to M, is depicted by the black arrows on top\nof the caps. One observes from Fig. 1(b) that the tickle field stron gly depends on the external\nstatic magnetic field orientation αand vanishes for α= 0◦andα= 90◦. Thus, for Bu= 0, the\nacoustic FMR signal will vanish for α=n·90◦,n∈Z, resulting in a four-fold symmetry. This\nis in stark contrast to conventional FMR, in which the RF magnetic fie ld is applied externally\nand does notdepend on α. Taken together, the above model suggests that the characte ristic\ndependence on the orientation of the external static magnetic fie ld can be used as a fingerprint\nto distinguish acoustically driven FMR from conventional FMR.\nWe next discuss our experimental results. Figure 2(a) shows the m agnitude of the trans-\nmission|S21|measured in the nickel/lithium niobate hybrid as a function of frequen cy at zero\n4external magnetic field. Several RFtransmission maxima, corresp onding to propagatingSAWs,\nare observed. The first transmission maximum occurs at the SAW de lay line fundamental fre-\nquencyof172MHz. SAWtransmissionisalsoobservedatoddharmon ics, whichallowschoosing\nthe magnetoelastic interaction frequency in a range from 172MHz t o 3.6GHz. Here, we con-\nstrain our discussion to the fundamental frequency as well as the 5th, 9thand 13thharmonic\nfrequency, as these frequencies exhibit the most intense SAW tra nsmission for the chosen IDT\nmetallization ratio of 0.5. Using appropriate IDT designs [17] it is possib le to excite SAWs\nat several almost arbitrarily chosen frequencies, or even a broad frequency range. The inset\nin Fig. 2(a) exemplarily shows the frequency dependence of |S21|around the 9thharmonic\n(1.55GHz) for two different values of the external magnetic field. The series of transmission\nlobes is characteristic of a SAW delay line passband [17]. As µ0|H|is changed from 100mT\nto 7mT the damping of the SAW increases by approximately 5dB. Figur e 2(b) depicts the\nevolution of the SAW transmission as a function of magnetic field stre ngth for a full magnetic\nfield sweep from −150mT to +150mT and back to −150mT. The data correspond to the SAW\ntransmission averaged over the FWHM of the central passband lob e|S21|. In addition to the\nhysteretic magnetization switching at |µ0H|<3mT already reported [13, 14], two absorption\nmaxima can be discerned as pronounced dips in Fig. 2(b). The latter s how no hysteresis, are\nsymmetric to zero external magnetic field, and are bothpresent in the magnetic field up- as well\nas the downsweep. Furthermore, the maximal SAW attenuation inc reases with frequency from\nless than 0 .1dB at 0 .17GHz to approximately 16dB at 2 .24GHz, and simultaneously shifts to\nlargerH. Due to the characteristic non-hysteretic behavior of the atten uation maxima together\nwith their shift as a function of frequency, we attribute the atten uation maxima to FMR. The\nslight asymmetry in the attenuation maxima in Fig. 2(b) is observed fo r allHorientations.\nFurther work will be necessary to determine its origin.\nTo identify the mechanism driving the FMR, a detailed study of the ang ular dependence\nis mandatory, as discussed in the context of Fig. 1(b). We thus per formed a series of mea-\nsurements to determine |S21|as a function of magnetic field magnitude and orientation. The\norientation αof the in-plane static magnetic field was varied in 5◦steps from −90◦to 90◦.\nFigure 3(a) shows the |S21|2data obtained as a false color plot. The color code represents\nthe normalized transmission |S21|2\nnorm=|S21|2(µ0H)/|S21|2(−150mT), with black indicating\nmaximal absorption. Figure 3(b) depicts the corresponding signal phase arg( S212), setting\narg(S212)(µ0H=−150mT) = 90◦, with black and yellow visualizing deviations of the signal\n5FIG. 3. (color online) Evolution of the SAW transmission as a function of the magnitude and orien-\ntation of the external magnetic field. Column (a) shows the ex perimentally determined |S21|2\nnorm. In\nresonance, the SAW transmission is attenuated, as evidence d by the black color. Column (b) shows\nthe experimentally determined arg( S212). Column (c) and column (d) show the transmitted SAW\npower density, simulated using Eq. (5). The simulation repr oduces all the salient features observed in\nthe experiment and demonstrates that the FMR is elastically driven.\nphase from 90◦(orange). As evident from Fig. 3, the SAW transmission exhibits a ch arac-\nteristic, approximately four-fold angular symmetry (four transm ission minima as a function of\n0◦≤α≤360◦for a given external magnetic field magnitude). Thus, in addition to t he shift of\nthe resonance field with frequency already discussed in the contex t of Fig. 2, the anisotropy of\nthe SAW magneto-transmission exhibits the characteristic fingerp rint of SAW-driven acoustic\nFMR. In particular, no attenuation occurs for α= 0◦andα=±90◦, as expected according to\nthe picture shown in Fig. 1(b). The fact that the attenuation maxim a are not located exactly\nat±45◦is due to a uniaxial in-plane magnetic anisotropy Bu∝ne}ationslash= 0, as discussed in the following.\nTo model the SAW magneto-transmission, we use a Landau-Lifshitz -Gilbert (LLG) ap-\nproach, taking into account RF magnetoelastic effects. We neglect spatial variations of the\nstrains and restrict the calculations to the collective FMR mode. The LLG equation [21, 22]\n∂tm=γm×µ0Heff+am×∂tm, (3)\ndescribes theevolution ofthemagnetizationdirection minaneffective magnetic field Heff, with\nγandabeing the gyromagnetic ratio and the Gilbert damping parameter, re spectively. We\nsolve theLLGinacartesiancoordinatesystem with thefirst axispoin ting alongtheequilibrium\ndirectionof m, whichinourcaseisinthefilmplaneatanangle θ0tothexaxisandthethirdaxis\n6pointing out-of-plane. In the following, the subscripts {1,2,3}refer to this frame of reference.\nThus,m={1,m2,m3}, wherem2andm3denote small deviations from equilibrium. The\nnon-vanishing component of the effective RF magnetic field h(t) =−∇mFRF(t)|θ=θ0[19, 20] is\ngiven by\nµ0h2(t) =−2B1ε(t)cos(θ0)sin(θ0). (4)\nIn contrast to classical FMR, where an external RF magnetic field is exploited to drive the\nresonance, Eq. 4 shows that acoustic FMR is excited by a purely inte rnal RF magnetic field\nh(t). The latter is due to RF spin mechanics, i.e., it is generated by the mag netoelastic\ninteraction of the SAW elastic strain field with the ferromagnet. This coupling between mand\nthe SAW results in a resonant attenuation and phase shift of the SA W transmission when the\ncondition for ferromagnetic resonance is met. We write the transm itted RF power density Pt\nas\nPt∝P0−1\n2µ0ωh†↔χh, (5)\nwhere the Polder susceptibility↔χis a function of the static magnetic free energy density F0,P0\nis the power density transmitted off resonance and ω= 2πνis the SAW angular frequency. The\nnon-vanishing components of the Polder susceptibility tensor read asχ22=Msγµ0ω1/D,χ33=\nMsγµ0ω2/Dandχ23=χ∗\n32=Msγµ0iω/DwithD=ω1ω2−ω2,ω1=γ(µ0H1+2Bd−2Buu2\n1)+\niωaandω2=γ(µ0H1+2Bu(u2\n2−u2\n1)) +iωa. Eq. (5) corresponds to the conventional FMR\nformula [8], but employs the purely virtual h(t) defined in Eq. (4) instead of an externally\napplied, real RF magnetic field.\nFigure 3(c) shows |Pt|norm=|Pt|(µ0H)/|Pt|(−150mT) calculated using Eq. (5), with Bd=\n+300mT, a= 0.25,ε= 10−5andBu= +4mT along the x-direction for the three frequencies\ninvestigated in the experiment. Figure 3(d) depicts the correspon ding transmission phases\narg(Pt). The simulation reproduces all the salient features observed in th e experiment, in\nparticular the angular dependence. Considering our rather simple m odel, the agreement is\nexcellent. This demonstrates that the SAW absorption is indeed cau sed by acoustically driven\nferromagneticresonance, i.e., byRFmagnetoelasticinteractions, andnotbyconventional FMR.\nMoreover, it corroborates the use of a virtual, purely internal h(t) on the same footing as a real\nexternally applied magnetic field in Eq. (5) a posteriori. We note that o ur damping constant\na= 0.25 is about a factor of ten larger than that expected from cavity F MR experiments [23].\nThis value was chosen to account for line broadening presumably due to the inhomogeneous\n7tickle field periodic with the SAW wavelength λ≈1.5µm at 2.24GHz, corresponding to a\nwavevector kSAW≈4×106m−1. In coplanar waveguide FMR such non-uniform excitation\nfields are known to lead to a line broadening by 100% already for k= 6×104m−1[24]. The\nmuch larger wavevectors in our case will therefore lead to consider ably higher linewidths, i.e.\na= 0.25. To circumvent such inhomogeneous broadening effects, one ma y miniaturize the\nferromagnetic thin film to lateral dimensions much smaller than the SA W wavelength. In such\nsamples, a determination of the influence of magnon-phonon intera ctions on magnetic damping\nshould be possible.\nIn conclusion, we have consistently found both in experiment and in s imulation that the\nmagnetoelastic interaction of a SAW with a ferromagnetic thin film allow s to excite FMR in\nthe film. The FMR is driven acoustically in the sense that no external R F magnetic field is\napplied tothe ferromagnet. Rather, a purely internal RFmagnetic field arises dueto magnetoe-\nlastic coupling between the SAW elastic strain field and the ferromagn et. Using a free energy\napproach as well as LLG calculations, we showed that the magnitude of the ”acoustic” RF\nmagnetic field characteristically depends on the orientation of the m agnetization. The angular\ndependence of the SAW transmission thus is a fingerprint of acoust ic FMR, as observed in our\nexperiments. Our experimental findings open a third alternative fo r the excitation of FMR,\nin addition to externally applied RF magnetic fields [25, 26] or the spin to rque effect [27–31].\nMore fundamentally, acoustic FMR can provide a pathway for the st udy of magnon-phonon\ninteraction phenomena studied predominantly on theoretical grou nds to date [7–10], or also\nfor mechanical spin pumping [32, 33]. The coupling between elastic and magnetic degrees of\nfreedom will open additional channels for spin dephasing, so that th e magnetization damping\nis linked to the magnon-phonon coupling strength. Furthermore, a detailed study of the in-\nteraction between magnonic and phononic dispersions, of the stre ngth of the coupling, of the\nanticrossing of these branches and of the generation of strongly coupled elastic/magnetic states\nare appealing challenges for the future.\nWe gratefully acknowledge stimulating discussions with B. Botters, R . Huber and D.\nGrundler. This work is financially supported by the DFG via project no s GO 944/3-1, SFB\n631 C3, and by the Cluster of Excellence Nanosystems Initiative Mun ich (NIM).\n∗goennenwein@wmi.badw.de\n8[1] H. Zheng et al., Science 303, 661 (2004).\n[2] N. A. Spaldin and M. Fiebig, Science 309, 391 (2005).\n[3] W. Eerenstein et al., Nat. Mater. 6, 348 (2007).\n[4] C. Bihler et al., Phys. Rev. B 78, 045203 (2008).\n[5] A. W. Rushforth et al., Phys. Rev. B 78, 085314 (2008).\n[6] M. Weiler et al., New J. Phys. 11, 013021 (2009).\n[7] H. F. Tiersten, J. Math. Phys. 5, 1298 (1964).\n[8] T. Kobayashi et al., Phys. Rev. B 7, 3273 (1973).\n[9] T. Kobayashi, R. C. Barker, and A. Yelon, Phys. Rev. B 7, 3286 (1973).\n[10] P. A. Fedders, Phys. Rev. B 9, 3835 (1974).\n[11] H. B¨ ommel and K. Dransfeld, Phys. Rev. Lett. 3, 83 (1959).\n[12] A. K. Ganguly, J. Appl. Phys. 47, 2696 (1976).\n[13] I. Feng et al., J. Appl. Phys. 53, 177 (1982).\n[14] R. F. Wiegert and M. Levy, J. Appl. Phys. 61, 4270 (1987).\n[15] R. F. Wiegert and M. Levy, IEEE Trans. Magn. 37, 2708 (2001).\n[16] R. F. Wiegert, J. Appl. Phys. 91, 8231 (2002).\n[17] S. Datta, Surface Acoustic Wave Devices (Prentice Hall, Englewood Cliffs, 1986).\n[18] S. Chikazumi, Physics of Ferromagnetism (Oxford Univ. Press, Oxford, 1997), 2nd ed.\n[19] J. Fidler and T. Schrefl, J. Phys. D: Appl. Phys. 33, R135 (2000).\n[20] M. J. Pechan et al., J. Appl. Phys. 89, 7514 (2001).\n[21] L. Landau and E. M. Lifshitz, Phys. Z. Sowjet. 8, 153 (1935).\n[22] T. Gilbert, IEEE Trans. Magn. 40, 3443 (2004).\n[23] S. M. Bhagat, L. L. Hirst, and J. R. Anderson, J. Appl. Phy s.37, 194 (1966).\n[24] G. Counil et al., J. Appl. Phys. 95, 5646 (2004).\n[25] C. Kittel, Phys. Rev. 73, 155 (1948).\n[26] L. R. Bickford, Phys. Rev. 78, 449 (1950).\n[27] J. C. Slonczewski, J. Magn. Magn. Mater. 159, L1 (1996).\n[28] E. B. Myers et al., Science 285, 867 (1999).\n[29] J. A. Katine et al., Phys. Rev. Lett. 84, 3149 (2000).\n[30] M. D. Stiles and A. Zangwill, Phys. Rev. B 66, 014407 (2002).\n[31] S. Zhang, P. M. Levy, and A. Fert, Phys. Rev. Lett. 88, 236601 (2002).\n9[32] A. A. Kovalev, G. E. W. Bauer, and A. Brataas, Phys. Rev. L ett.94, 167201 (2005).\n[33] O. Mosendz et al., Phys. Rev. Lett. 104, 046601 (2010).\n10" }, { "title": "1710.05166v1.Repulsive_polarons_in_alkaline_earth__like__atoms_across_an_orbital_Feshbach_resonance.pdf", "content": "Repulsive polarons in alkaline-earth(-like) atoms across an orbital Feshbach resonance\nTian-Shu Deng,1, 2Zhuo-Cheng Lu,3Yue-Ran Shi,3Jin-Ge Chen,3Wei Zhang,3, 4,\u0003and Wei Yi1, 2,y\n1Key Laboratory of Quantum Information, University of Science and Technology of China,\nChinese Academy of Sciences, Hefei, Anhui, 230026, China\n2Synergetic Innovation Center of Quantum Information and Quantum Physics,\nUniversity of Science and Technology of China, Hefei, Anhui 230026, China\n3Department of Physics, Renmin University of China, Beijing 100872, China\n4Beijing Key Laboratory of Opto-electronic Functional Materials and Micro-nano Devices,\nRenmin University of China, Beijing 100872, China\n(Dated: June 9, 2021)\nWe characterize properties of the so-called repulsive polaron across the recently discovered orbital\nFeshbach resonance in alkaline-earth(-like) atoms. Being a metastable quasiparticle excitation at\nthe positive energy, the repulsive polaron is induced by the interaction between an impurity atom\nand a Fermi sea. By analyzing in detail the energy, the polaron residue, the e\u000bective mass, and\nthe decay rate of the repulsive polaron, we reveal interesting features that are intimately related to\nthe two-channel nature of the orbital Feshbach resonance. In particular, we \fnd that the life time\nof the repulsive polaron is non-monotonic in the Zeeman-\feld detuning bewteen the two channels,\nand has a maximum on the BEC-side of the resonance. Further, by considering the stability of a\nmixture of the impurity and the majority atoms against phase separation, we show that the itin-\nerant ferromagnetism may exist near the orbital Feshbach resonance at appropriate densities. Our\nresults can be readily probed experimentally, and have interesting implications for the observation\nof itinerant ferromagnetism near an orbital Feshbach resonance.\nI. INTRODUCTION\nThe recently discovered orbital Feshbach resonance\n(OFR) in173Yb opens up the avenue of investigating\nstrongly-interacting many-body physics using alkaline-\nearth(-like) atoms [1{3]. In an OFR, the spin-exchange\ninteraction between the ground1S0and the long-lived ex-\ncited3P0hyper\fne manifolds can be tuned by an exter-\nnal magnetic \feld. It follows that the wealth of precision\nquantum control techniques, which have been developed\nfor the purpose of quantum metrology and quantum in-\nformation using the clock-state manifolds ( f1S0;3P0g),\ncan be employed to engineer highly non-trivial many-\nbody scenarios [4{31]. Recent studies in this regard range\nfrom interaction-induced topological states [32, 33], to\nimpurity problems such as the Kondo e\u000bects [34{39] and\nthe polaron to molecule transitions [40, 41]. Naturally,\nthe key properties of these phenomena are \frmly based\non the features of interactions of an OFR.\nLike the interactions of Feshbach resonance in alkali\natoms, the interactions of OFR can be understood as\nthe resonant scattering between an open and a closed\nchannel. Consider two alkaline-earth(-like) atoms respec-\ntively in the1S0(denoted asjgi) and the3P0(denoted\nasjei) manifolds, as J= 0 for these so-called clock-state\nmanifolds, the nuclear and the electronic spin degrees of\nfreedom are decoupled. Denoting a particular nuclear\nspin statemI(mI+1) in each manifold as j \"i (j #i),\nwe may associate the open channel with the jg#iand\n\u0003Electronic address: wzhangl@ruc.edu.cn\nyElectronic address: wyiz@ustc.edu.cnje\"istates, and the closed channel with the jg\"iand\nje#istates. Due to the di\u000berential Zeeman shift in the\nclock-state manifolds [42, 43], an external magnetic \feld\ncan conveniently shift the detuning between the open-\nand the closed-channel scattering thresholds. Further, as\nthe short-range interaction of the OFR occurs either in\nthe electronic spin-singlet and nuclear spin-triplet chan-\nnel, or the electronic spin-triplet and nuclear spin-singlet\nchannel, it couples the closed- and the open channels to-\ngether. The scattering resonance occurs when the en-\nergy of a bound state in the closed channel is tuned\nto the open-channel scattering threshold. As a result\nof OFR, a crossover from the Bardeen-Cooper-Schrie\u000ber\n(BCS) to the Bose-Einstein condensation (BEC) regime\ncan be realized in alkaline-earth(-like) atoms by tuning\nthe magnetic \feld, which is similar to the magnetic Fesh-\nbach resonance in alkali atom. However, the existence of\nmultiple nuclear spin states, as well as the spin-exchange\ninteractions in the OFR complicate the two-body scatter-\ning process, and lead to rich physics in the many-body\nsetting.\nAn illuminating example here is the system consist-\ning of a mobile impurity interacting with its environ-\nment. As the limiting case of a many-body system in\nthe large polarization limit, mobile impurity and its as-\nsociated quasi-particle excitations contain valuable in-\nformation of the underlying system. Whereas impurity\nproblems in the background of Bose gases or Fermi con-\ndensates have attracted considerable attention in recent\nyears [44{63], here we focus on the case of an impurity\nagainst a non-interacting Fermi sea. In alkali atoms,\nit has been shown that the impurity can either form\na tightly bound molecule with a majority atom, or in-\nduce collective particle-hole excitations in the Fermi seaarXiv:1710.05166v1 [cond-mat.quant-gas] 14 Oct 20172\nand form the so-called Fermi polaron [64{69]. A polaron\nto molecule transition has been observed experimentally,\nas the interaction is tuned. Further, at positive ener-\ngies, a so-called repulsive polaron branch exists, which\nis metastable and associated with the elusive itinerant\nferromagnetism [68{75]. In OFR, a recent theoretical\nstudy suggests that the transition between the attrac-\ntive polaron and the molecule also exists when tuning\nthe magnetic \feld [40, 41]. However, the existence and\nproperties of the repulsive polaron branch have not been\ninvestigated.\nIn this work, we characterize properties of the repulsive\npolaron across the OFR, using the parameters of173Yb\natoms as a concrete example. As illustrated in Fig. 1, we\nconsider a single impurity atom in the je\"istate, which\ninteracts with a Fermi sea of atoms in the jg#istate.\nWhile the impurity and the background atoms are ini-\ntially in the open channel, the spin-exchange interactions\nwould scatter atoms into the closed channel. Adopting\nthe T-matrix formalism [72, 73], we demonstrate the ex-\nistence of a metastable repulsive polaron branch at pos-\nitive energies across the OFR. We characterize various\nproperties of the repulsive polaron, such as the energy,\nthe polaron residue, the e\u000bective mass, and the decay\nrate. In particular, we identify unique features in all of\nthese quantities, which are intimately related to the two-\nchannel nature of the OFR. An interesting result of the\ninter-channel scattering is that the life time of the repul-\nsive polaron is non-monotonic in the e\u000bective interaction\nstrength, and has a maximum on the BEC-side of the\nresonance. We further analyze the condition for the ex-\nistence of itinerant ferromagnetism in these atoms near\nan OFR. By considering the stability of a homogeneous\nmixture of the impurity je;\"iatoms and the majority\njg;#iatoms against phase separation, we show that a\nphase-separated state, and hence the itinerant ferromag-\nnetism, can be stabilized beyond a critical Zeeman-\feld\ndetuning. Since such a conclusion is conditional on the\nstability of the repulsive polaron, we further demonstrate\nthat for appropriate atomic densities, a parameter win-\ndow exists where the system favors phase separation and\nthe repulsive polaron is long-lived and away from the\nmolecule-hole continuum. Our \fndings can be readily\nprobed experimentally, and have interesting implications\nfor the observation of itinerant ferromagnetism near an\nOFR.\nThe paper is organized as follows. In Sec. II, we present\nthe T-matrix formalism for Fermi polarons in the context\nof an OFR. We demonstrate the existence of the repulsive\npolaron, and characterize its energy by calculating the\nspectral function in Sec. III. We then study in detail the\npolaron residue and the e\u000bective mass in Sec. IV, where\nkinks in these properties are identi\fed and associated\nwith resonant scatterings in the many body background.\nWe characterize the decay rate of the repulsive polaron in\nSec. V, and discuss in detail the potential stability region\nof the itinerant ferromagnetism near an OFR in Sec. VI.\nFinally, we summarize in Sec. VII.II. T-MATRIX FORMALISM\nWe start from the non-interacting Hamiltonian corre-\nsponding to the con\fguration in Fig. 1(a)\nH0=X\nk\u000fo\nk(ay\ng;#kag;#k+ay\ne;\";kae;\";k)\n+X\nk\u000fc\nk(ay\ne;#kae;#k+ay\ng;\"kag;\"k); (1)\nwhereay\nj;\u001b;k(aj;\u001b;k) creates (annihilates) an atom in the\ncorresponding pseudo-spin state jj;\u001bi(j2fg;eg,\u001b2\nf\";#g) with momentum k. Here,\u000fo\nk=~2k2=2mand\n\u000fc\nk=~2k2=2m+\u000e=2. The detuning between the two\nchannels\u000e\u0011\u0001g\u0000\u0001e= (gg\u0000ge)\u0016BBoriginates from\nthe di\u000berential Zeeman shift of the clock states in the\npresence of a magnetic \feld B, wheregg(ge) is the Lande\ng-factor for thejgi(jei) manifold, and \u0016Bis the Bohr\nmagneton.\nThe typical inter-orbital spin-exchange interaction of\nan OFR can be written as\nHint=g+\n2X\nqAy\n+(q)A+(q) +g\u0000\n2X\nqAy\n\u0000(q)A\u0000(q);(2)\nwhere we have\nA\u0006(q) =X\nk(ae;#;kag;\";q\u0000k\u0007ae;\";kag;#;q\u0000k);(3)\nand the interaction strengths g\u0006are related to the\nphysical ones via the renormalization relation 1 =g\u0006=\n1=~g\u0006\u0000P\nk1=2\u000fo\nkwith ~g\u0006= 4\u0019~2a\u0006=m. Throughout\nthis work, we adopt the parameters of173Yb atoms, with\na+= 1900a0anda\u0000= 219:5a0[2, 3, 76, 77].\nDiagrammatically, the polaron properties can be cal-\nculated using the retarded self-energy of the impurity\natom [72], which is given by [see Fig. 1(b)]\n\u0006(Q;E) =Zdq\n(2\u0019)3Zd!\n2\u0019G0\ng#(q;!)Too(q+Q;E+!);\n(4)\nwhereG0\ng;#(q;!) = (!+i0+\u0000\u000fo\nq)\u00001is the free-fermion\npropagator of the majority atoms, and Toois the T-\nmatrix describing the open-channel scattering processes.\nHereEandQare respectively the energy and the center-\nof-mass momentum of the self-energy, and !is the Mat-\nsubara frequency. Due to the spin-exchange nature of the\ninteraction, the open- and the closed-channel scattering\nmatrices are coupled. Accordingly, there should be four\nkinds of T-matrices Too,Toc,Tco, andTcc, with the in-\ncoming and the outgoing states being in either the open\nor the closed channel, as indicated by the superscript la-\nbels. As discussed in Ref. [31], under the ladder approx-\nimation, we may write down a set of coupled equations\nfor the T-matrices, which lead to the solution\nToo(q;!) =1\n2(g++g\u0000)\u0000g+g\u0000\u001fc\n1\u00001\n2(g++g\u0000)(\u001fo+\u001fc) +g+g\u0000\u001fo\u001fc:(5)3\n|e,↓t\n|g,↓t(a)\nToo(b)\n=Gg↓G0\nGe↑0Ge↑0|e,↑t\n|g,↑t∆e\n∆g\nȞGe↑0Ge↑0\nGe↑0Ge↑0\nFIG. 1: (a) Level diagram of an OFR in alkaline-earth(-\nlike) atoms. An impurity of je;\"iis immersed in a majority\nFermi sea ofjg;#iatoms, and can be scattered to the other\ntwo atomic states forming the closed channel via interaction.\n\u0001g=gg\u0016BBand \u0001 e=ge\u0016BBare the Zeeman shifts of the\njgiandjeimanifolds, respectively. (b) The one-hole polaron\nself-energy \u0006 near an OFR. The solid lines with arrows indi-\ncate free propagator G0forjg;#iorje;\"i, and the square Too\nindicates the T-matrix with the incoming and the outgoing\nstates being both in the open channel.\nHere, the pair propagators for the closed and the open\nchannel\u001fc(q;!) and\u001fo(q;!) can be written as\n\u001fc(q;!) =X\nk1\n!+i0+\u0000\u000fc\nk\u0000\u000fc\nq\u0000k; (6)\n\u001fo(q;!) =X\njkj>kF1\n!+i0+\u0000\u000fo\nk\u0000\u000fo\nq\u0000k; (7)\nwhere the Fermi wave vector kFis related to the Fermi\nenergyEFofjg;#iatoms asEF=~2k2\nF=2m. From the\nequations above, we see that \u001fo(q;!) and\u001fc(q;!) are\nisotropic in q. For the convenience of discussion, we de-\n\fneq=jqj.\nSubstituting Eq. (5) into (4), we obtain\n\u0006(Q;E) =X\nq0 corresponds to the repulsive polaron.\nIn contrast to the attractive polaron, which is undamped\nunder the ladder approximation here, the repulsive po-\nlaron, being a mestastable quasipaticle excitation with\nE+>0, features a \fnite width in the spectral func-\ntion as illustrated in Fig. 3, which originates from the\ndecay into low-lying states. Under the inter-orbital spin-\nexchange interactions of the OFR, as we will show later,\nthe \fnite spectral width and hence the decay of the repul-\nsive polaron mainly come from the resonant coupling of\nthe quasi-particle excitation to the open- and the closed-\nchannel scattering continuum. Finally, we notice the ex-\nistence of a broad wing between the two polaron peaks as\nshown in Fig 3, which corresponds to the molecule-hole\ncontinuum.\nIV. IMPURITY RESIDUE AND THE\nEFFECTIVE MASS\nWe now characterize the impurity residue and the ef-\nfective mass of the repulsive polaron. For a polaron ex-\ncitation, the quasi-particle residue is de\fned as [72]\nZ\u0006=1\n1\u0000Reh\n@\u0006(0;!)\n@!i\f\f\f\f\f\n!=E\u0006; (12)\nand its e\u000bective mass as\nm\u0003\n\u0006=1\nZ\u00061\n1 + Reh\n@\u0006(Q;!)\n@Q2i\f\f\f\f\f\nQ=0;!=E\u0006; (13)4\n-2 02\n/E0-3-2-1012E/E0\n246810\nFIG. 2: False color plot of the spectral function A(Q= 0;E)\nof an impurityje;\"iin a Fermi sea of non-interacting jg#i\nparticles on the \u000e{Eplane. The solid red lines depict the\npolaron energies given by Eq. (11), and the dashed green\nline is the molecular energy. The light-blue area between the\ntwo polaronic branches is the molecule-hole continuum. The\nupper repulsive polaron branch merges into the molecule-hole\ncontinuum for \u000e&3:8E0. Here, we de\fne the unit of energy\nE0=~2k2\n0=2m, where the unit Fermi wave vector k3\n0= 6\u00192n0\nand the unit density n0= 5\u00021013cm\u00003. In this plot, we take\nn=n0.\nwhere the subscript + ( \u0000) labels the repulsive (attrac-\ntive) branch of polarons.\nWe have shown the quasi-particle residue as well as\nthe e\u000bective mass of the repulsive polaron in Fig. 4. For\ncomparison, we have also plotted the residue and the ef-\nfective mass of the attractive polaron. In an OFR and\nunder the setup illustrated in Fig. 1, the resonance occurs\nat\u000e0\u00183:06E0, and the system is on the BCS-side of the\nresonance for \u000e>\u000e 0. In Fig. 4, we see that as \u000eincreases\n(i.e., moves towards the BCS side of the resonance), Z+\ndecreases and m\u0003\n+increases, which are qualitatively con-\nsistent with the case of alkali atoms. A prominent dif-\nference in the current case is the existence of kinks in\nboth the residue and the e\u000bective mass at \u000e=E+and\n\u000e=E++EF=2. The occurrence of these kinks can be\nexplained by the qualitative di\u000berence, between regions\nwith di\u000berent values of \u000e, in the way that the atoms in\nthe open-channel Fermi sea are scattered into the closed-\nchannel continuum in forming the repulsive polaron.\nThe location of the kinks can be determined analyti-\ncally by considering the scattering process between the\nimpurity and the majority atoms, in which the out-going\nstates are at the closed-channel scattering threshold. In\nparticular, because \u000erepresents the closed-channel de-\ntuning of the two atoms and E+is the interaction-\ninduced energy shift of the impurity atom, at \u000e=E+an\nimpurity atom with zero momentum can interact with\na majority atom at the bottom of the Fermi sea (with\nq= 0), which are resonantly scattered to two atoms\nin the closed-channel scattering threshold. Likewise, at\n-2 02\nE/E00510A(0,E)=0(a)\n-2 02\nE/E00510A(0,E)=1(b)\n-2 02\nE/E00510A(0,E)=2(c)\n-2 02\nE/E00510A(0,E)=3(d)FIG. 3: The spectral function A(Q= 0;E) as functions of E\nwith di\u000berent detunings. We have taken the same parameters\nas those in Fig. 2.\n\u000e=E++EF=2, an impurity atom with zero momentum\ninteracts with a majority atom on the Fermi surface (with\nq=kF), which, under the momentum conservation, are\nresonantly scattered to two atoms in the closed channel\neach with a momentum q=kF=2. The process can there-\nfore be qualitatively described as resonant scatterings in\nthe many-body background.\nTo further demonstrate this point, in Fig. 5(a) and\n5(b), we explicitly show the imaginary parts of the pair\npropagators in the open- and the closed-channel, respec-\ntively, on the \u000e{qplane. As the imaginary parts of the\npair propagators are related to the removable singular-\nities in the summation of Eqs. (6) and (7), they re\rect\nthe contribution to the polaron self-energy as atoms in\nthe Fermi sea with momentum qE ++EF=2, on the other hand,\n\u001fc(q;E++\u000fq) can be completely real for any q < kF.\nHence, none of the atoms in the Fermi sea can be reso-\nnantly scattered into the closed-channel continuum, and\nthe particle-hole excitations in the repulsive polaron is\nopen-channel dominated. Therefore, in di\u000berent regions\nof\u000e, the closed-channel scattering continuum contribute5\n-2 02\n/E000.51Z(a)\n-2 02\n/E0-505m*/m(b)\nFIG. 4: (a) Quasi-particle residues Z\u0006for the attractive (blue\ndashed) and repulsive (red solid) polarons as functions of \u000e.\n(b) E\u000bective masses of the attractive (blue dashed) and the\nrepulsive (red solid) polarons as functions of \u000e.\nin qualitatively di\u000berent ways to the polaron self-energy,\nwhich gives rise to the appearance of kinks at the bound-\naries of these regions.\nV. DECAY RATE OF THE REPULSIVE\nPOLARON\nBeing a mestable state, the repulsive polaron can de-\ncay into low-lying states. Experimentally, it has been\nshown that for alkali atoms, the dominating decay chan-\nnel for the repulsive polaron is the coupling to the bare\nimpurity state in the attractive-polaron branch, so long\nas the interaction is not in the deep BEC regime [75].\nWe assume that the case with alkaline-earth(-like) atoms\nis similar. One should then include the corresponding\ndecay channel in the diagrams leading to the repulsive\npolaron self-energy. Such a decay rate can be calculated\nas [75]\n\u0000 =\u00002Z+[Im\u0006(0;E+)]; (14)\nwhereZ+is the residue for the repulsive polaron. Fur-\nther, we can replace the free-fermion propagator G0\ne\"with\n(1\u0000Z+)G0\ne\"in the self-energy \u0006, which implies substitut-\ning\u001fowith (1\u0000Z+)\u001fo. This leads to \u0000 =P\njqj Teff\nC,\nM1andM2are exchange decoupled and their orienta-\ntions can be changed independently by applying a suit-\nableexternalmagneticfield, H. Thus, avariationintem-\nperature and/or field can produce switching between the\nparallel (P) and antiparallel (AP) mutual orientations of\nM1andM2in the system.8,9\nThe key element in the F 1/f/F2sandwich described\nabove (the so-called Curie-switch orCurie-valve ) is the\nweakly ferromagnetic spacer, f, which should have a nar-\nrow ferromagnetic-to-paramagnetic transition and have\ntheTeff\nCvalue tunable in fabrication. Diluted ferromag-\nnetic alloys, such as Ni-Cu, is the natural choice for the\nspacer material, since the Curie temperature of bulk10as\nwell as film11–14samples of Ni xCu100−xalloys depends\nalmost linearly on Ni concentration.\nThe experiments described in Refs. 8 and 9 confirmed\nthe concept of temperature-controlled P to AP switch-\ning in nanostructures F 1/f/F2, in particular containing aNixCu100−x(x= 35÷72 at. %) spacer enclosed by an\nexchange-pinned Co 90Fe10layer and a free Ni 80Fe20(Py)\nlayer: Py/Ni xCu100−x/Co90Fe10/Mn80Ir20(hereinafter –\nF1/f/F2/AF, AF denoting antiferromagnetic Mn 80Ir20).\nSince the earlier work primarily aimed at understanding\nthe switching effect itself, little attention was paid to the\neffect of the spacer-mediated exchange on the ferromag-\nnetic resonance in the structure.\nThis work investigates the magnetic resonance prop-\nerties of the Curie-switch, experimentally and theoreti-\ncally, aiming at understanding the mechanisms involved\nandobtainingtheintrinsicphysicalparametersgoverning\nthe spin dynamics in the system.\nII. THEORY\nConsider an F 1/f/F2/AF multilayer outlined above,\nwhere a weakly ferromagnetic spacer (f) is sandwiched\nbetweenmagneticallysoft(F 1)andhard(F 2)layers,with\nF2exchange pinned by an antiferromagnetic layer (AF).\nThe thicknesses of F 1, F2and f are l1,l2, andd, respec-\ntively.\nOur calculations of the magnetic resonance fields will\nassume that F 1and F 2are single domain. For small\nlayer thicknesses and strong exchange interaction, in the\nweak excitation limit typical of ferromagnetic resonance\n(FMR) experiments, this assumption is well justified.15\nSpacer f with magnetization mprovides a relativelyweak\ncouplingbetweentheouterferromagnets,F 1andF2. The\nanalysis aims to determine the effect of this interlayer\nexchange coupling, variable in strength as a function of\ntemperature, on the FMR.\nWe use the classical Landau-Lifshitz approach16,17to\ndescribetheF 1/f/F2/AFmultilayerandfocusonthecase\nwherethe quasistaticexternalfield Handthealternating\nfieldhareinthe film plane. Axes OxandOyaredirected2\nFIG. 1. (Color online) Schematic of F 1/f/F2/AF multilayer\nand chosen coordinate system in FMR measurements.\nalongHandh, respectively (Fig. 1). With Ozperpen-\ndicular to the film plane, the magnetization vectors of F 1\nand F2can be expressed as\nMi=Mi(sinθicosϕi,sinθisinϕi,cosθi),(1)\nwherei= 1,2;Miis the saturation magnetization of i–\nth layer; θiandϕiare the polar and azimuthal angles,\nrespectively.\nThe exchange bias between F 2and AF can be mod-\nelled using an effective biasing field Hbacting only on\nmagnetization M2.18It will be shown below that for fully\ndescribing the FMR-effects of interest in this work, it is\nsufficient to consider only two cases, for which the exter-\nnal fieldHis either parallel or antiparallel to Hb. Corre-\nspondingly, the value of Hbcan be either positive ( φ= 0\nin Fig. 1) or negative ( φ=π) along the biasing axis.\nThe magnetic energy of the i-th ferromagnetic layer in\nthe above geometrical notations is\nWi=Sliwi, (2a)\nwi= 2πM2\nicos2θi−MiHicosϕisinθi\n−Mihsinϕisinθi,(2b)\nwhereSis the areaof the film surface, wiis thei-th layer\nenergy density, H1=H, andH2=H+Hb. The first\nterms in Eq. (2b) originates from demagnetizing energy,\nwhile the second and third terms describe the energies\nof the interaction of the layers’ magnetizations with the\nquasi-static Hand alternating hexternalmagneticfields,\nrespectively.\nTo simplify the equations, let us recall that in the case\nofathin film, its highout-of-planedemagnetizationfieldspreventthe magnetization vectorfrom stronglydeviating\nfrom the xOyplane. In this case, θican be represented\nasθi=π/2+εi, where|εi| ≪1.\nIn the small-signal approximation relevant for FMR,\nthe magnetizationvectorsofF 1andF2arenearlyaligned\nwith the Oxaxis (the easy axis, also the direction of ex-\nternal field H) and perform only weak oscillations near\nthe ground state under the microwave excitation h. This\nmeans that |ϕi| ≪1. The limits of validity of this ap-\nproximationwill be discussed in the experimental section\nbelow.\nIn the above small-signal thin-film approximation, the\nmagnetic energy density of the system becomes\nwi=−MiHi+(4πM2\ni+MiHi)ε2\ni/2\n+MiHiϕ2\ni/2−Mihϕi.(3)\nF1and F2are exchange coupled through a weakly fer-\nromagnetic spacer f. The case where the spacer is highly\nmagnetically diluted and nominally (in the bulk) is para-\nmagnetic was consideredin Ref. 9, with the interlayerex-\nchange mediated via induced proximity ferromagnetism.\nHere we consider the case where the spacer is diluted\nsuch that it is nominally ferromagnetic and can mediate\ndirect exchange between the outer ferromagnetic layers,\nwith the exchange coupling strength being a steep func-\ntion of temperature near the Curie point of the spacer.\nWedenotethetemperaturedependent saturationmag-\nnetization of the spacer as m. Assuming again that the\nmagnetization is uniform in the xOyplane, the spacer\nenergy density can be written as\nw=αm2\n2/bracketleftigg/parenleftbigg∂θ\n∂z/parenrightbigg2\n+sin2θ/parenleftbigg∂ϕ\n∂z/parenrightbigg2/bracketrightigg\n−mHm\n2−Hmcosϕsinθ−hmsinθsinϕ ,(4)\nwhereαis the constant of exchange interaction, θand\nϕare polar and azimuthal angles, respectively, of the\nspacer magnetic moment m,Hmis the magnetostatic\nfield in the system.\nThe value of the magnetostatic field can be easily\nderived from Maxwell’s equation: div B= div(Hm+\n4πm) = 0. Since both the magnetization and therefore\nmagnetostatic field depend only on one special variable,\nz, the magnetostatic field becomes: Hm=−4πmzez=\n−4πmcosθez, where ezis the unit vector along the z\naxis.\nTaking into account Eq. (4), the Landau-Lifshitz equa-\ntions in the angular form become\n∂2θ\n∂ξ2=−d2\nΛ2/bracketleftbigg\nsinθcosθ+sinθ∂ϕ\n∂τ+H\n4πmcosθcosϕ+h\n4πmcosθsinϕ/bracketrightbigg\n, (5a)3\n∂\n∂ξ/parenleftbigg\nsin2θ∂ϕ\n∂ξ/parenrightbigg\n=d2\nΛ2/bracketleftbigg\nsinθ∂ϕ\n∂τ+H\n4πmsinθsinϕ−h\n4πmsinθcosϕ/bracketrightbigg\n. (5b)\nThe new dimensionless variables, normalized to the char-\nacteristic length and time in the problem, introduced in\nEqs. (5a) and (5b) are ξ=z/dandτ= 4πtγm, wheret\nis the time and γis the gyromagnetic ratio. Λ =/radicalbig\nα/4π\nis the magnetic exchange length.19\nIf the spacer thickness dis much smaller than the\nmagnetic exchange length Λ ( d≪Λ), the right side in\nEqs. (5a) and (5b) becomes a small correction, which in\nthe first approximation can be neglected.\nAs a result, only the exchange terms survive:\n∂2θ\n∂ξ2= 0, (6a)\n∂\n∂ξ/parenleftbigg\nsin2θ∂ϕ\n∂ξ/parenrightbigg\n= 0. (6b)\nThe solution, which satisfies the requirement of con-\ntinuity of the polar and azimuthal components at the\ninterfaces between the layers, has the form:\nϕ(z) =ϕ2+(ϕ1−ϕ2)z/d,\nε(z) =ε2+(ε1−ε2)z/d, (7)\n0≤z≤d.The resulting magnetic energy of the spacer is\nW=SJ/bracketleftbig\n(ϕ1−ϕ2)2+(ε1−ε2)2/bracketrightbig\n/2,(8)\nwhereJ=αm2/d= 4πΛ2m2/d.\nTo determine the resonance conditions for the layered\nsystem under consideration, we express the Lagrange\nfunction in terms of the angle:\nL=2/summationdisplay\ni=1/parenleftbigg\n−SliMi\nγεi∂ϕi\n∂t−Wi/parenrightbigg\n+J/bracketleftbig\n(ϕ1−ϕ2)2+(ε1−ε2)2/bracketrightbig\n/2.(9)\nThe variational equations following from Eg. (9) are\nequivalent to the Landau-Lifshitz equations:\n\niHωH+h1 0 −h1\n4πM1+H+h1−iHω −h1 0\n0 −h2 iHω H+Hb+h2\n−h2 0 4πM2+H+Hb+h2−iHω\n×\nε1\nϕ1\nε2\nϕ2\n=\nh\n0\nh\n0\n (10)\nwherehi=J/Mili=αm2/dMili= 4πΛ2m2/dMili.\nHere,Hω=ω/γ,ω= 2πf,γis the gyromagnetic ra-\ntio,hiis the characteristic field of exchange interaction\nbetween the layers.\nThe characteristic fields of the resonance modes ofthe collective spin dynamics in the system are found by\nequating the determinant of matrix (10) to zero. This\nresults in two branches in the functional form of Hω(H).\nThe first resonance branch, corresponding to the reso-\nnance field of F 1, has the form:\nH2\nω= (Hr1+h1)(4πM1+Hr1+h1)−h1h2/bracketleftbigg\n1−(2πM1+Hr1)(2πM2+Hr1)\n2πHr1(M2−M1)/bracketrightbigg\n−h1h2Hb\nHr14π[π(M2+M1)+Hr1](M2−M1)−(2πM1+Hr1)(2πM2+Hr1)\n[2π(M2−M1)]2,(11)\nwhereHr1is the external field producing FMR in F 1[see\nFig.2(a)]. Onlytermsofordernothigherthanquadratic\ninkiwere kept in Eq. (11).Thevalueof Hr1dependsonwhethertheexternalmag-\nnetic field is parallel ( ↑↑) or antiparallel ( ↑↓) to the ex-\nchangebiasfield Hb. Itiseasytoshowthatthedifference\nin the resonance fields, ∆ Hr1=H↑↓\nr1−H↑↑\nr1, has the form:4\n∆Hr1=H↑↓\nr1−H↑↑\nr1=h1h2Hb\nH0\nr14π/bracketleftbig\nπ(M2+M1)+H0\nr1/bracketrightbig\n(M2−M1)−(2πM1+H0\nr1)(2πM2+H0\nr1)\n(2πM1+H0\nr1)[2π(M2−M1)]2,(12)\nwhereH0\nr1= (H↑↓\nr1+H↑↑\nr1)/2 =/radicalbig\n(2πM1)2+H2ω−\n2πM1−h1.\nIt follows from Eq. (12) that ∆ Hr1is proportional to a\nproduct of h1h2. This means that ∆ Hr1sharply changes\nin the vicinity of the Curie point of the spacer as a result\nof the sharp increase in mat the para-to-ferromagnetic\ntransition (see Eq. (10). Expectedly, ∆ Hr1goes to zero\nasTincreasesabovetheCuriepointofthe spacer. Inthis\nhigh-Tlimit, thereisnocouplingbetweenF 1andF2, and\nEq. (11) describes the resonance field of the decoupled\nsoft outer ferromagnet F 1.\nTo find the resonance fields for F 2, we keep only terms\nof the order not higher than linear in hi. The results for\nH↑↓\nr2andH↑↑\nr2are\nH↑↑\nr2=/radicalbig\n(2πM2)2+H2ω−2πM2−h2−Hb,\nH↑↓\nr2=/radicalbig\n(2πM2)2+H2ω−2πM2−h2+Hb.(13)\nAgain, sharp changes in H↑↑\nr2andH↑↓\nr2are expected in\nthe vicinity of the Curie point of the spacer. At high\ntemperatures where h2→0, the difference between H↑↓\nr2\nandH↑↑\nr2naturally becomes 2 Hb.\nIt follows from Eq. (13) that for sufficiently high values\nofM2andh2, theH↑↑\nr2branchcanfallintonegativefields,\nwhere it cannot be observed experimentally.\nIII. EXPERIMENT\nA. Samples and measurements\nThe experiments were carried out on two sets of\nmultilayered samples, in which either the spacer thick-\nness,d, or its composition, x, were varied. The\nfirst set was Py(10 nm)/Ni 54Cu46(d)/Co90Fe10(5 nm)/\nMn80Ir20(12 nm) (hereinafter – F 1/Ni54Cu46(d)/F2/AF)\nwith the spacer thicknesses d= 3, 4.5, and 6 nm. The\nsecond set was Py(10 nm)/Ni xCu100−x(6 nm)/Co 90Fe10\n(5nm)/Mn 80Ir20(12nm)(hereinafter��F 1/NixCu100−x(6\nnm)/F 2/AF), with x= 54, 62 and 70 at.%. The multi-\nlayers were deposited at room temperature on thermally\noxidized silicon substrates using magnetron sputtering in\nan AJA Orion 8-target system. The exchange pinning\nbetween the ferromagnetic Co 90Fe10and antiferromag-\nnetic Mn 80Ir20was set in during deposition of the mul-\ntilayers using an in-plane magnetic field Hdep≈1 kOe.\nOther fabrication details are similar to those described\nin Refs. 8 and 9.\nIn addition to the multilayers, single-layer Py (10 nm)\nand Co 90Fe10(5 nm) films were prepared under identi-\ncal technological conditions. FMR measurements on the\nsingle-layer films were carried out to extract the mag-\nnetizations of Py and Co 90Fe10layers and use them forsubsequent multilayer-FMR modelling and characteriza-\ntion [e.g., using Eqs. (12) and (13)].\nThe FMR measurements were performed using an X-\nband ELEXSYS E500 spectrometer equipped with an\nautomatic goniometer. The operating frequency was\nf= 9.44 GHz. FMR spectra for various in-plane dc-\nfield angles were studied in the temperature range of 120\nto 400 K.\nB. Results and discussion\n1. Measured FMR spectra\nFig. 2 (a) shows two typical FMR spectra for a\nF1/f/F2/AF multilayer, for which the external magnetic\nfield is parallel (solid line) or antiparallel (red dashed\nline) to the exchange bias field Hb(T= 300 K). The res-\nonancesignalsfrombothF 1andF2layersareclearlyvisi-\nble and are separated in field. As expected [see Eqs. (12)\nand (13)], the resonance conditions for both layers de-\npend on the mutual orientation of HandHb[Fig. 2(b)].\nConsistent with the predicted behavior of Eq. (13), the\nH↑↑\nr2branch for F 2extrapolates into negative fields [see\nFig.2(b)]. Intheremainderofthepaperwethereforedis-\ncuss and in-depth analyzeonly the H↑↓\nr2resonancebranch\nas regards the dynamics of the pinned F 2layer.\nTo understand the details of how the interlayer ex-\nchange coupling affects the spin dynamics of the free\nlayer (F 1), the angular dependence of Hr1was studied\nat various temperatures for various spacer thicknesses.\nThe typical data shown in Fig. 3 indicate that the po-\nsition of the resonance peak is angle-dependent and this\nangular asymmetry becomes strongeras the temperature\nis lowered. At the same time, the Hr1vsφdependence\nbecomes more pronounced as the spacer thickness de-\ncreases. For all the cases shown in Figs. 2 (b) and 3,\neachHr1(φ) data set is well fitted using a model charac-\nteristicofathin filmwith unidirectionalanisotropy(solid\nlines in Fig. 2 (b) and Fig. 3).\nA more detailed analysis shows that there is an addi-\ntional small contribution from uniaxial anisotropy. The\nextracted uniaxial anisotropy field ( ∼4 Oe) is weakly de-\npendent on temperature and spacer composition. This\ncontribution potentially originates from some ordered\nconfiguration at NiCu/Py interface formed during the\nmultilayer deposition in a magnetic field Hdep(used for\nexchange pinning F 2).5\n/s32 /s32/s32/s61/s32/s48 /s40\n/s98/s41\n/s32 /s32/s32/s61/s32/s49/s56/s48 /s40\n/s98/s41\n/s48 /s53/s48/s48 /s49/s48/s48/s48 /s49/s53/s48/s48/s70\n/s49/s47/s78/s105\n/s53/s52/s67/s117\n/s52/s54/s40/s52/s46/s53/s32/s110/s109 /s41/s47 /s70\n/s50/s47/s65/s70\n/s72\n/s114/s50/s72\n/s114/s49\n/s40/s98/s41/s40/s97/s41/s84 /s61/s51/s48/s48/s32/s75\n/s32/s32/s100/s80/s47/s100/s72 /s32/s32/s40/s97/s46/s117/s46/s41\n/s72 /s32/s40/s79/s101/s41/s49/s50/s48/s48 /s49/s50/s53/s48 /s49/s51/s48/s48 /s49/s51/s53/s48/s72\n/s114/s49/s32/s70\n/s49/s32/s32/s40/s80/s121/s41\n/s32/s32\n/s72 /s32/s40/s79/s101/s41\n/s48 /s57/s48 /s49/s56/s48 /s50/s55/s48 /s51/s54/s48/s48/s53/s48/s48/s49/s48/s48/s48/s49/s53/s48/s48\n/s84 /s61/s51/s48/s48/s32/s75\n/s32/s72\n/s114/s50/s32/s40/s70\n/s50/s32/s61/s32/s67/s111\n/s57/s48/s70/s101\n/s49/s48/s32/s41\n/s32/s72\n/s114/s49/s32/s40/s70\n/s49/s32/s61/s32/s80/s121/s41/s72\n/s114/s32/s32/s40/s79/s101/s41\n/s32/s40/s100/s101/s103/s46/s41\nFIG. 2. (Color online) (a) FMR spectra for F 1/Ni54Cu46(4.5\nnm)f/F 2/AF for parallel (solid line) and antiparallel (red\ndashed line) orientations of the external magnetic field H\nwith respect to the exchange biasing field Hb. The upper\ninset shows an enlarged view of the signal from F 1. (b) The\ndependence of the resonance fields of F 1and F 2on the angle\nφbetween HandHbextracted from data sets such as those\nillustrated in (a).\n2. FMR modelling procedure\nExpressions(12) and (13) were used for calculating the\ntemperature dependence of the resonance field asymme-\ntry ∆Hr1=H↑↓\nr1−H↑↑\nr1for F1(Py), and of the resonance\nfieldH↑↓\nr2for F2(Co90Fe10). These expressions contain\nthe values of the saturation magnetization M1andM2\nfor the Py and Co 90Fe10layers, respectively, which are\ntemperature dependent. The M1(T) andM2(T) depen-\ndences, used in the data analysis to follow, are shown in\nFig. 4 and were obtained from the FMR data taken on\nsingle-layer Py (10 nm) and Co 90Fe10(5 nm) films pre-\npared under the same conditions as the multilayers. The\nKittel’s formulas for isotropic thin films20,21were used to\ncalculate the M1(T) andM2(T) shown.\nThe key quantity determining the behavior of ∆ Hr1\nandH↑↓\nr1, is the spacer magnetization m, averaged over\nthe layerthickness [seeEq. (10)]. The spaceris ferromag-\nnetic below the Curie point and nominally paramagnetic/s56/s48/s48/s57/s48/s48/s49/s48/s48/s48/s49/s49/s48/s48/s49/s50/s48/s48\n/s32/s40/s97/s41/s100/s32/s61/s32/s51/s32/s110/s109/s72\n/s114/s49/s32/s40/s79/s101/s41\n/s32/s49/s50/s48/s32/s75/s32\n/s32/s50/s57/s48/s32/s75/s70\n/s49/s47/s78/s105\n/s53/s52/s67/s117\n/s52/s54/s40/s100 /s41/s47/s70\n/s50/s47/s65/s70\n/s45/s49/s56/s48 /s45/s57/s48 /s48 /s57/s48 /s49/s56/s48/s49/s49/s48/s48/s49/s50/s48/s48/s49/s51/s48/s48/s49/s52/s48/s48\n/s32/s49/s50/s48/s32/s75/s32\n/s32/s50/s57/s48/s32/s75/s100/s32/s61/s32/s54/s32/s110/s109\n/s40/s98/s41/s72\n/s114/s49/s32/s40/s79/s101/s41\n/s32/s32/s40/s100/s101/s103/s46/s41\nFIG. 3. (Color online) Angular dependences of the resonance\nfields of the soft layer (F 1) in F 1/Ni54Cu46(d)/F2/AF with\nthe spacer thickness d= 3 nm (a) and 6 nm (b).\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48 /s52/s48/s48/s52/s48/s48/s54/s48/s48/s49/s52/s48/s48/s49/s54/s48/s48\n/s32/s32 /s80/s121 \n/s32/s32 /s67/s111\n/s57/s48/s70/s101\n/s49/s48/s77 /s32/s32/s40/s101/s109/s117/s47/s99/s109/s51\n/s41\n/s84 /s32/s32/s40/s75/s41\nFIG. 4. (Color online) Temperature dependences of the mag-\nnetizations of Py and Co 90Fe10layers obtained from the FMR\ndata on the respective single-layered films.\naboveit. It has previouslybeen shown, however,that the\nproximity effect at the interface with a strong ferromag-\nnet induces noticeable magnetization in a paramagnetic\nor weakly ferromagnetic metal and may give rise to an\nincrease in its Curie point.9,22,23The proximity length\nis an order of magnitude greater than the atomic spac-\ning and for the case where the spacer thickness dis of6\nthe order of a few nanometers, the induced magnetiza-\ntion penetrates through the spacer thickness.8,9For this\nreason, to account for the proximity effect in our calcula-\ntions, it was assumed that (i) there is an additional field\nHproxwhich acts on the spacer magnetization mand (ii)\nfor the spacer sandwiched between strong ferromagnets\nF1and F 2, the Curie point Teff\nCdiffers from that in the\nbulk.\nThem(T,Hprox) dependence due to the prox-\nimity effect was modelled using the mean-field\napproximation,15,24wherethe magnetizationis described\nby the Brillouin function:\nm\nm0=Bj(m) =ξ1cosh/bracketleftbigg\nξ1/parenleftbigg\nζ1m\nT+ζ2Hprox\nT/parenrightbigg/bracketrightbigg\n−ξ2cosh/bracketleftbigg\nξ2/parenleftbigg\nζ1m\nT+ζ2Hprox\nT/parenrightbigg/bracketrightbigg\n;\nξ1=2j+1\n2j;ξ2=1\n2j;ζ1=3j\nj+1Teff\nC\nm0;ζ2=νm0; (14)\nwherejis the total angular momentum per Ni xCu100−x\nformula unit, m0is the saturation magnetization at\nT= 0 K,Hproxis the effective field reflecting the prox-\nimity effect at the interfaces with the strong ferromag-\nnets F 1and F 2, andTeff\nCis the effective Curie tempera-\nture. Coefficient νequalsµ/(ρNA), where µandρare\nthe molar mass and density of Ni xCu100−x, respectively,\nNAis the Avogadro constant, and kBis the Boltzmann\nconstant.15,24\nThe coefficient νfor our Ni xCu100−xalloy was esti-\nmated to be near 8 .2×10−8cm3K erg−1forxin the\nvicinity of 60 at.%. The initial values of jandm0were\nchosen based on the data calculated in Ref. 8 for bulk\nNixCu100−x, and the value of jwas kept fixed through-\nout the analysis. Since the magnetizationand Curie tem-\nperature of the Ni xCu100−xspacer are expected to differ\nfrom those in the bulk, specifically due to the proximity\neffect,m0was chosen as one of the variable parameters\nin fitting the experimental data.\nThe proximity effect is expected to be most pro-\nnounced in the vicinity of Teff\nC. Theab-initio calculations\nof this effect for F 1/NixCu100−x(x,d)/F2/AF atT∼Teff\nC\nwere detailed in Ref. 9. Based on a comparison of the\nvalues for the average magnetic moment /angbracketleftm/angbracketrightobtained in\nRef. 9 and the mcalc(T) obtained using Eq. 14, it was\nfound that mcalcatT∼Teff\nCis approximately equal to\n/angbracketleftm/angbracketrightforHprox≈100 kOe. This value of Hproxwas kept\nfixed in all subsequent calculations.\nAnother important quantity affecting the spin dy-\nnamics in the system is the exchange bias field\nHb. Based on the magnetometry measurements on\nF1/NixCu100−x(x,d)/F2/AF reported in Ref. 8, Hbwas\nobtained for a range of xanddvalues (for 300 K). These\nand the additional data reported in Ref. 9 make it pos-\nsible to conclude that for our samples with x >52,Hb\nis only weakly temperature dependent. The calculation\ntherefore assumed Hb(T) = const. The specific fixed jandHbvalues used in the calculations, among other pa-\nrameters and variables, are presented in Table I.\nSummarizing, the variable parameters used to fit the\ntheoretical ∆ Hr1(T) andH↑↓\nr2(T) to the experimental\ndata were the effective Curie temperature ( Teff\nC) and sat-\nuration magnetization at T= 0 (m0) of the Ni xCu100−x\nspacer, as well as the characteristic magnetic exchange\nlength (Λ). It will be shown below that for the case un-\nder study, the resulting values of Λ are about two times\ngreater than the spacer thickness d. This is within the\nlimits of the approximation d≪Λ used in the analysis.\n3. FMR data analysis\nFigs. 5 (a) and (b) show the temperature dependences\nofH↑↓\nr1−H↑↑\nr1andH↑↓\nr2for F1/Ni54Cu46(d)/F2/AF sam-\nples obtained from the measured FMR spectra as well as\nthe respective theoretical fits using the above analysis.\nA good agreement between the experiment and theory is\nobtained for realistic values of the fitting parameters.\nThe temperature dependence of the spacer magneti-\nzationm/m0obtained from fitting the resonance fields\nis shown in Fig. 5 (c) for different values of the spacer\nthickness. The proximity of the strongly ferromagnetic\nlayers F 1and F 2has essentially no effect on the low-\ntemperature magnetization of the spacer but is the dom-\ninant factor determining its effective Curie point Teff\nC.\nThe changes in Teff\nCstrongly depend on the spacer thick-\nness: the smaller the dand, therefore, the stronger the\nproximity effect of the interfaces, the higher the Teff\nC.\nFig. 6 (a, b) shows the measured resonance fields\nH↑↓\nr1−H↑↑\nr1andH↑↓\nr2as a function of temperature\nfor F1/NixCu100−x(6 nm)/F 2/AF with different Ni-\nconcentration in the spacer, fitted to theory using\nEqs. (12) and (13). The agreement is good, includ-\ning the case of the highest Ni-concentration with non-\nmonotonous temperature dependence of the resonance\nfield asymmetry (70 at.% Ni in Fig. 6 (a)).\nThe dependence of the spacer magnetization on tem-\nperature extracted from fitting the data in Figs. 6 (a, b)\nis shown in Fig. 6 (c) for different Ni-concentration of the\nspacer. It is clear in this case that the proximity of the\nstrongly ferromagnetic outer layers affects both the low-\ntemperature magnetization m0and the effective Curie\npointTeff\nCof the spacer – the greater the x, the higher\nthem0andTeff\nC.\nHaving performed the data analysis, it is now informa-\ntive to discuss the accuracyof the theoretical assumption\nmade in Section 2 in describing the spin dynamics in an\nF1/f/F2/AF spin-thermionic system. One key assump-\ntion was that the magnetization vectors of the F 1and\nF2layers are parallel to the external field Hand perform\nonly weak oscillation under the influence of the alternat-\ningfieldcomponent h. Thisassumptionisstrictlycorrect\nfor the case where His parallel to Hb, but the opposite\nantiparallelcase( H↑↓Hb)mustbeconsideredwithcare.7\nTABLE I. Magnetic parameters of F 1/NixCu100−x(x,d)/F2/AF multilayers.\nx(at.% Ni) d(nm) jaHb(Oe)bTeff\nC(K) m0(emu/cm3) Λ (nm)\n0.54 3.0 0.19 180 450 120 11 ±2\n0.54 4.5 0.19 240 320 120 11 ±2\n0.54 6.0 0.19 300 250 120 11 ±2\n0.62 6.0 0.21 280 350 140 12 ±2\n0.70 6.0 0.23 210 550 150 13 ±2\naValues calculated from data of Ref. 9 under assumption that L andeg-factor equals 2.\nbValues obtained from magnetic hysretesys loops at room temp erature of Ref. 8.\n/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48\n/s32/s100/s32 /s61/s32/s51/s32/s110/s109\n/s32/s100 /s32/s61/s32/s52/s46/s53/s32/s110/s109\n/s32/s100 /s32/s61/s32/s54/s32/s110/s109\n/s32/s32/s72\n/s114/s49/s32 /s32/s72\n/s114/s49/s32/s32/s40/s79/s101/s41\n/s50/s48/s48/s52/s48/s48/s54/s48/s48/s56/s48/s48/s32\n/s40/s99/s41/s40/s97/s41/s72\n/s114/s50/s32/s32/s40/s79/s101/s41\n/s32/s32/s70\n/s49/s47/s78/s105\n/s53/s52/s67/s117\n/s52/s54/s40/s100 /s41/s47/s70\n/s50/s47/s65/s70\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48 /s52/s48/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s32/s100 /s32/s61/s32/s51/s32/s110/s109\n/s52/s46/s53/s32/s110/s109\n/s54/s32/s110/s109/s40/s98/s41/s109 /s47/s109\n/s48/s32/s32/s40/s97/s46/s117/s46/s41\n/s32/s32\n/s84 /s32/s32/s40/s75/s41\nFIG.5. (Coloronline)Temperaturedependencesof H↑↓\nr1−H↑↑\nr1\n(a) andH↑↓\nr2(b) for F 1/Ni54Cu46(d)/F2/AF samples for three\nspacerthicknesses(symbols). Boldsolid linesshowtheore tical\nH↑↓\nr1−H↑↑\nr1andH↑↓\nr2vsTdependences fitted to the measured\ndata using Eqs. (12) and (13), respectively. (c) Temperatur e\ndependence of the normalized spacer magnetization m/m0\nobtained using the above fitting of the resonance fields.\nThevaluesof HbforF1/NixCu100−x(x,d)/F2/AFmul-\ntilayers are listed in Table I. As follows from the M−H\ndata presented in Refs. 8 and 9, the magnetization of\nthe F1/f/F2/AF multilayer fully saturates when the ap-\nplied reversing field H↑↓exceeds (1.2 ÷1.5) times Hb.\nThus, the above assumption should be valid when both\nresonance fields, H↑↓\nr1andH↑↓\nr2, are higher than (1.2 ÷/s48/s49/s48/s48/s50/s48/s48/s51/s48/s48\n/s120 /s32/s61/s32/s53/s52/s32/s97/s116/s46/s37/s120 /s32/s61/s32/s54/s50/s32/s97/s116/s46/s37/s120/s32 /s61/s32/s55/s48/s32/s97/s116/s46/s37/s32/s120 /s61/s55/s48/s32/s97/s116/s46/s37\n/s32/s120 /s61/s54/s50/s32/s97/s116/s46/s37\n/s32/s120 /s61/s53/s52/s32/s97/s116/s46/s37/s72\n/s114/s49/s32 /s32/s72\n/s114/s49/s32/s32/s40/s79/s101/s41\n/s32/s32\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48 /s52/s48/s48/s48/s46/s48/s48/s46/s53/s49/s46/s48/s40/s98/s41\n/s32/s32/s109 /s47/s109\n/s48/s32/s40/s97/s46/s117/s46/s41\n/s84 /s32/s40/s75/s41/s70\n/s49/s47/s78/s105\n/s120/s67/s117\n/s49/s48/s48/s45 /s120/s40/s54/s32/s110/s109/s41/s47/s70\n/s50/s47/s65/s70\n/s50/s48/s48/s52/s48/s48/s54/s48/s48/s56/s48/s48/s32\n/s40/s99/s41/s40/s97/s41/s72\n/s114/s50/s32/s32/s40/s79/s101/s41\n/s32/s32\nFIG.6. (Coloronline)Temperaturedependencesof H↑↓\nr1−H↑↑\nr1\n(a) and H↑↓\nr2(b) for F 1/NixCu100−x(6 nm)/F 2/AF samples\nfor three Ni concentrations (symbols). Bold solid lines sho w\ntheoretical H↑↓\nr1−H↑↑\nr1andH↑↓\nr2vsTdependences fitted to\nthe measured data using Eqs. (12) and (13), respectively. (c )\nTemperature dependences of the normalized spacer magneti-\nzationm/m0obtained usingtheabovefittingoftheresonance\nfields.\n1.5)Hb. For the samples under study, H↑↓\nr1is greater\nthan 900 Oe, so the first condition, H↑↓\nr1>(1.2÷1.5)Hb,\nis well satisfied. The data in Figs. 5 and 6 indicate that\nthe second condition, H↑↓\nr2>(1.2÷1.5)Hb, is also well\nsatisfied for temperatures above ∼180 K.\nThe developed approach, combining theory and ex-8\n/s53/s52 /s54/s50 /s55/s48/s50/s48/s48/s51/s48/s48/s52/s48/s48/s53/s48/s48/s54/s48/s48\n/s32/s32 /s70\n/s49/s47/s78/s105\n/s120/s67/s117\n/s49/s48/s48/s45 /s120/s40/s54/s32/s110/s109/s41 /s47/s70\n/s50/s47/s65/s70/s84/s101/s102/s102\n/s67/s32/s32/s40/s75/s41\n/s120 /s32/s32/s40/s97/s116/s46/s37/s32/s78/s105/s41/s51/s46/s48 /s52/s46/s53 /s54/s46/s48\n/s32/s32/s32 /s70\n/s49/s47/s78/s105\n/s53/s52/s67/s117\n/s52/s54/s40/s100 /s41/s47/s70\n/s50/s47/s65/s70\n/s32/s100 /s32/s40/s110/s109/s41/s32\nFIG. 7. (Color online) DependenceoftheeffectiveCurie poin t\nTeff\nCof the spacer Ni xCu100−xon its thickness d(at fixed com-\nposition x= 54 at.% Ni) and Ni content x(at fixed thickness\nd= 6 nm).\nperiment, makes it possible to extract and analyze the\nthickness and composition dependence of the character-\nistic parameters of the spacer, which are summarized in\nTable I.\nFig. 7 presents the model-extracted dependence of the\neffective Curie point Teff\nCof the spacer on its thickness\ndand Ni content x. We conclude that small variations\nin bothdandxresult in significant variations in Teff\nC.\nWe also note that the values of the Curie temperature\nof NixCu100−xsandwiched between strong ferromagnets\nare much greater than the corresponding values in the\nbulk form of this diluted magnetic alloy – ( Tbulk\nC≈120\nK and 300 K for x= 54 and 70 at.% Ni, respectively)10.IV. CONCLUSION\nFerromagnetic resonance properties of F 1/f/F2/AF\nmultilayers, where spacer f has a low Curie point com-\npared to strongly ferromagnetic F 1and F 2, have been\nanalyzed theoretically and investigated experimentally.\nThe spacer-mediated exchange coupling is shown to\nstrongly affect the resonance fields of both F 1and F 2\nlayers. It is found that the key magnetic parameters of\nthe spacer which govern the magnetic resonance in the\nsystem are the magnetic exchange length (Λ), the effec-\ntive saturation magnetization at T= 0 (m0), and the\neffective Curie temperature ( Teff\nC) of the spacer.\nBy theoretically fitting the measured FMR data, the\nvalues of Λ, m0, andTeff\nCare obtained for the technolog-\nically significant ranges in Ni-concentration ( x= 54÷70\nat.% Ni) and thickness ( d= 3÷6 nm) of the spacer.\nThe values thus obtained are entirely different from the\ncorresponding quantities in the bulk. The developed ap-\nproach to spin dynamics in the system enables such de-\ntailed quantitative characterization that is otherwise is\ndifficult or impossible obtain in terms of direct measure-\nments due to the built-in strong proximity effect.\nTheinferredmagnetisminthekeyelementofthestruc-\nture – the spacer, acting as an interlayer exchange-spring\n– shows a great sensitivity and thereby high tunability\nof its properties versus the degree of magnetic dilution,\ngeometry, and temperature. These results should be use-\nful for designing high-speed nanodevices based on spin-\nthermionic control.\nACKNOWLEDGMENTS\nSupport from the Swedish Stiftelse Olle Engkvist\nBuggm¨ astare,SwedishResearchCouncil(VRgrant2014-\n4548), the Science and Technology Center in Ukraine\n(project P646), and the National Academy of Sciences\nof Ukraine (projects 0115U003536 and 0115U00974) are\ngratefully acknowledged.\n∗anatolii@kth.se; Correspondence author\n1B. Dieny, V. S. Speriosu, S. S. P. Parkin, B. A. Gurney, D.\nR. Wilhoit, and D. Mauri, Phys. Rev. B 43, 1297 (1991).\n2Nanomagnetism and Spintronics edited by by T. Shinjo\n(Elsevier, Amsterdam, 2009).\n3A. Hirohata and K. Takanashi, J. Phys. D: Appl. Phys.\n47, 193001 (2014).\n4S. Andersson and V. Korenivski, IEEE Trans. Magn. 46,\n2140 (2010).\n5S. Andersson and V. Korenivski, J. Appl. Phys. 107,\n09D711 (2010).\n6A. M. Kadigrobov, S. Andersson, D. Radi´ c, R. I. Shekhter,\nM. Jonson, and V. Korenivski, J. Appl. Phys. 107, 123706\n(2010).7A.M. Kadigrobov, S.Andersson, HeeChulPark, D.Radi´ c,\nR. I. Shekhter, M. Jonson, and V. Korenivski, J. Appl.\nPhys.111, 044315 (2012).\n8A. F. Kravets, A. N. Timoshevskii, B. Z. Yanchitsky, M.\nA. Bergmann, J. Buhler, S. Andersson, and V. Korenivski,\nPhys. Rev. B 86, 214413 (2012).\n9A. F. Kravets, Yu. I. Dzhezherya, A. I. Tovstolytkin, I. M.\nKozak, A. Gryshchuk, Yu. O. Savina, V. A. Pashchenko,\nS. L. Gnatchenko, B. Koop, and V. Korenivski, Phys. Rev.\nB90, 104427 (2014).\n10D. J. Chakrabarti, D. E. Laughlin, S. W. Chen, and Y.\nA. Chang, in Phase Diagrams of Binary Copper Alloys\nedited by P. Subramanian, D. Chakrabarti, and D. Laugh-\nlin (ASM International, Materials Park, OH, 1994), p. 276.9\n11A. P. Thakoor and K. L. Chopra, J. Appl. Phys. 48, 3850\n(1977).\n12I.Bakonyi, E.T´ oth-K´ ad´ ar, J.T´ oth, T. Becsei, T.Tarn´ o czi,\nand P. Kamasa, J. Phys.: Condens. Matter 11, 963 (1999).\n13A. Rusanov, R. Boogaard, M. Hesselberth, H. Sellier, and\nJ. Aarts, Physica C 369, 300 (2002).\n14A. F. Kravets, A. N. Timoshevskii, B. Z. Yanchitsky, O.\nYu. Salyuk, S. O. Yablonovskii, S. Andersson, and V. Ko-\nrenivski, J. Magn. Magn. Mater. 324, 2131 (2012).\n15A. Aharoni, Introduction to the Theory of Ferromagnetism\n(Oxford University Press, Oxford, 1996).\n16L. D. Landau and E. M. Lifshitz, in The Collection of\nWorksedited by L. D. Landau (Nauka, Moscow, 1969)\n[in Russian], Vol. 1, p. 128.\n17C. Serpico, I. D. Mayergoyz, and G. Bertotti, J. Appl.\nPhys.93, 6909 (2003).18J. Nogues, J. Sort, V. Langlais, V. Skumryev, S. Surinach,\nJ. S. Munoz, and M. D. Baro, Phys. Rep. 422, 65 (2005).\n19Gavin S. Abo, Yang-Ki Hong, Jihoon Park, Jaejin Lee,\nWoncheol Lee, andByoung-ChulChoi, IEEE Trans. Magn.\n49, 4937 (2013).\n20Ch. Kittel, Phys. Rev. 73, 155 (1948).\n21A. G. Gurevich and G. A. Melkov, Magnetization Oscilla-\ntions and Waves (CRC Press, Boca Raton, FL, 1996).\n22N. Garcia and A. Hernando, J. Magn. Magn. Mater. 99,\nL12 (1991).\n23I. Navarro, M. Ortuno, and A. Hernando, Phys. Rev. B\n53, 11656 (1996).\n24S. Chikazumi, Physics of Ferromagnetism (Oxford Univer-\nsity Press, New York, 2005)." }, { "title": "2308.11996v1.Parallel_simulation_of_the_magnetic_moment_reversal_within_the__φ_0__Josephson_junction_model.pdf", "content": " 1 PARALLEL SIMULATION OF THE MAGNETIC MOMENT \nREVERSAL WITHIN THE 0-JOSEPHSON JUNCTION MODEL \nM. Bashashina,b,*, E. Zemlyanayaa,b, I. Rahmonova,b,c \naJoint Institute for Nuclear Research , Joliot -Curie str. 6, 141980 Dubna, Russia \nbDubna State University , University str. 19, 141980 Dubna, Russia \ne-mail: bashashinmv@jinr.ru \ncMoscow Institute of Physics and Technology , 14170 0, Dolgoprudny , Russia \nReceived \nAbstract — Periodic structure of the magnetization reversal domains is studied within the superconductor –\nferromagnetic –superconductor φ0-junction model. The model is described by the Cauchy problem for the system \nof nonlinear ordinary equations which is numerically s olved by means of the 2 -step Gauss –Legendre method. Two \nversions of parallel implementation on the basis of MPI and OpenMP techniques have been developed. Efficiency \nof both versions is confirmed by test calculations. An effect of frequency of ferromagneti c resonance on the \nconfiguration MR domains has been investigated. The calculations have been performed at the HybriLIT Platform \nof the JINR Multifunctional information and computing complex . \nINTRODUCTION \nIn the superconductor –ferromagnetic –superconductor structures, the spin -orbit coupling \nin ferromagnetic layer without inversion symmetry provides a mechanism for a direct (linear) \ncoupling between the magnetic moment and the superconducting cu rrent [1]. Such Josephson \njunctions are called φ0-junction. A perspective to control magnetic properties of φ0-junction s \nby means of the superconducting current as well as effect of magnetic dynamics on the \nsuperconducting current is of great interest from the viewpoint of practical applications in \nnanoelectronic devises [1-3]. One of the important directions in this field is a study of \nphenomenon of magnetization reversal (MR) in φ0-junction [4]. Different scenarios f or the MR \nrealization and the periodic structure of MR intervals are demonstrated in [5] . This contribution \naims to study an influence of frequency of ferromagnetic resonance on the configuration MR \ndomains. The basic formulae of the mathematical model are given . Numerical approach \nincluding the parallel implementation is briefly described . Results of numerical simulations of \nthe MR phenomenon depending on the parameters of model are presented . Also, the results of \ntest calculations demonstrating effect of parallel implementation on the basis of MPI and \nOpenMP techniques are given. 2 THEORETICAL MODEL \nWe use the theoretical framework as in [ 4]. The model of the superconductor –\nferromagnetic –superconductor φ0-Josephson junction is based on the Landau –Lifshitz –Gilbert \nequation which describes t he dynamics of the magnetization vector 𝑚⃗⃗ in ferromagnetic layer in \nthe φ 0-junction : \n𝑑𝑚⃗⃗ \n𝑑𝑡=−𝜔𝐹\n1+𝑚⃗⃗ 2𝛼2{[𝑚⃗⃗ ×𝐻⃗⃗ ]+𝛼[𝑚⃗⃗ (𝑚⃗⃗ 𝐻⃗⃗ )−𝐻⃗⃗ 𝑚⃗⃗ 2]}, (1) \nwhere α is damping parameter, ωF is normalized frequency of ferromagnetic resonance , 𝐻⃗⃗ is an \neffective magnetic field vector with the components \n{𝐻𝑥=0 \n𝐻𝑦=𝐺𝑟sin(𝜑(𝑡)−𝑟𝑚𝑦(𝑡))\n𝐻𝑧=𝑚𝑧(𝑡) , (2) \nG – relation of Josephson energy to energy of magnetic anisotropy, r – the spin -orbit coupling \nparameter , mx,y,z corresponds the x,y,z-component of mag netization vector 𝑚⃗⃗ . The Josephson \nphase difference φ can be found as follows: \n𝑑𝜑\n𝑑𝑡=𝐼𝑝𝑢𝑙𝑠𝑒(𝑡)−sin(𝜑−𝑟𝑚𝑦), (3) \nwhere the pulse current Ipulse is given by \n𝐼𝑝𝑢𝑙𝑠𝑒={𝐴𝑆, 𝑡∈[𝑡0−12∆𝑡,𝑡0+12∆𝑡]⁄ ⁄ \n0, otherwise. (4) \nHere As is the amplitude of the pulse current, and Δt is the time interval, in which the pulse \ncurrent is applied. The system of equations (1) with effective fiel d (2), (3) and with the pulse \ncurrent (4) describes the dynamics of the φ0-junction. The initial conditions are following : \n𝑚𝑥 (0)=0,𝑚𝑦 (0)=0,𝑚𝑧 (0)=1, 𝜑0=0. (5) \nProblem (1 -5) is a Cauchy problem for a system of ordinary differential equations . Numerical \nsolution is based on the implicit two -step Gauss –Legendre method which was shown in [6] to \nprovide more accurate calculations in comparison with t he explicit Runge –Kutta scheme . \nRESULTS OF NUMERICAL SIMULATIONS \nMagnetic reversal is an effect when the component mz(t) change s its initial value mz(0)=+1 and \nstabilizes on the value mz(Tmax)=–1 for sufficiently large T max. Numerical study is red uced to \nmassive numerical simulations the Cauchy problem (1-5) with varied parameters of model. In \n[5], the results of such simulations at the planes of parameters ( G, α) and ( G, r) are presented \nwhile all other parameters have the following values: at As = 1.5; r = 0.1; t0 =25; Δt = 6. \nSimulations were carried out with the time stepsize ht=0.01 while t≤Tmax=1000 . MR was 3 indicated using the condition |mz(Tmax)+1|<ε, where ε=0.0001. It was shown in [ 5] that the MR \nrealization is periodic when the value of G changes, i.e. there are intervals on G where MR \noccurs and intervals where MR is absent . The purpose of this work is to demonstrate a \ndependence of configuration of the MR realization intervals on the parameter ωF (normalized \nfrequency of ferromagnetic resonance). To this end, massive numerical simulations were \ncarried out on the (G, α) and ( G, r) planes in order to reveal the domains where the MR takes \nplace . Three values of ωF were used in calculations : ωF=0.1, 1, and 10. The results are presented \nin Figures 1 and 2. \nFigure 1 demonstrate s the MR domains at the (G, α)-plane. Colored areas correspond the \npairs of parameters G and α where the MR occur s. One sees that at small ωF=0.1 we obtained \na solid domain of G between 0 and 60 where the MR takes place. In case ωF=1, the MR domains \nare in the form of wide bands alternating with areas where the MR does not occur . As for the \ncase of large ωF=10 – the MR domains structure looks rather chaotic due to the very narrow \nalternating intervals of presence and absence of the magneti zation reversal. \n \nFig. 1. Magnetization reversal domains at the (G, α)-plane depending on value of ωF. The \ncalculations were carried out with the G-stepsize ΔG=0.1, α-stepsize Δα=0.001 . \n \nFig. 2. Magnetization reversal domains at (G,r)-plane depending on value of ωF. The \ncalculations were carried out with the G-stepsize Δ G = 0.1 , r-stepsize Δ r = 0.005 . \nFigure 2 shows the MR domains at the (G, r)-plane. As in Figure 1, one observes that the \nmagnetization reversal intervals (colored) alternate with the areas where the MR does not occur. \nIntervals of presence and absence of MR become narrow as ωF grows from 0.1 to 10. \n 4 EFFECT OF PARALLEL IMPLEMENTATI ON \nSince numerical investigation of the MR effect requires massive calculation s for numerical \nsolution of the C auchy problem (1 -5) at a large number of p airs of parameters at (G,α) and ( G,r) \nplanes , the parallel implementation have been developed to reduce the execution time and to \naccelerate the numerical study . In addition to the MPI parallel implementation presented in [7], \nwe developed the OpenMP version to expand the possibilities of numerical research on different \ncomputing systems . In both MPI and OpenMP ve rsions, t he paralleli sm is based on a \ndistribution of the points of (G,α) or (G,r) plane between parallel processes . The calculations \nwere made on two platforms of the Multifunctional Information Computing Centre [8] – the \nHybriLIT cluster and the Govorun supercomputer. The number P of parallel MPI processes and \nOpenMP threads was varied from 1 to 32. Both MPI and OpenMP implementation s are \nconfirmed to be quite effective and provide almost the same execution time in calculations at \nboth HybriLIT c luster and Govorun supercomputer. \nFigure 3 demonstrates that the execution time in test calculations at the (G,r) plane \ndecreases hyperbolically with increasing a number P of parallel MPI processes and OpenMP \nthreads. Weak oscillations of MPI -curves are explained by the strengthening of the interaction \nof parallel processes with increasing P (for details, see the recent work [9]). \n \nFig. 3. The execution time of MPI and OpenMP calculations in depend ence on P. \nSUMMARY \nMassive numerical simulations i n the wide range of parameters within the 0-junction \nmodel allowed to obtained domains where the magnetic moment is reversed. An influence of \nthe normalized frequency of the magnetic resonance ωF on the periodic structure of MR \ndomains has been studied. Test calculations confirm that both MPI and OpenMP parallel \nversions of the C++ computer code can provide the high-performance investigation of the MR \nphenomenon within the 0-junction model. \n 5 The work was supported by the Russian Science Foundation grant № 18-71-10095 \n(https://rscf.ru/project/21 -71-03009/ ). \nREFERENCES \n1. Buzdin A. I. Direct Coupling Between Magnetism and Superconducting Current in the \nJosephson φ0 Junction // Physical Review Letters . 2008 . V.101 ,107005 . \n2. Buzdin A. I . Proximity effects in superconductor -ferromagnet heterostructures // Rev. Mod. \nPhys. 2005 . V. 77 , 935. \n3. Konschelle F., Buzdin A. Magnetic Moment Manipulation by a Josephson Current // Phys. \nRev. Lett. 2009 , 102, 017001 . \n4. Shukrinov Yu. M., Rahmonov I. R., Sengupta K., Buzdin A. Magnetization reversal by \nsuperconducting current in Josephson junctions Appl. Phys. Lett. 2017 . 110, 182407 . \n5. Atanasova, P.K., Panayotova, S.A., Rahmonov, I.R. Shukrinov, Yu. M., Zemlyanaya E. V., \nBashashin, M. V., Periodicity in the Appearance of Intervals of the Reversal of the Magnetic \nMoment of a 0 Josephson Junction // JETP Letters . 2019 , 110(11):722 -726. \n6. Atanasova P. K., Panayotova S. A., Shukrinov Yu. M. and Rahmonov I. R. Numerical \nSimu lation of the Stiff System of Equations within the Spintronic Model // Lecture Notes \nin Computer Sciences . 2019 . 11189 , 301–308. \n7. Stefani Panayotova, Maxim Bashashin, Elena Zemlyanaya, Pavlina Atanasova, Yury \nShukrinov, Ilhom Rahmonov. Parallel Numerical Si mulation of the Magnetic Moment \nReversal within the ф0-Josephson Junction Spintronic Model. // European Physics Journal, \nWeb of Conferences . 2020 .Vol. 226, 02018 \n8. Adam Gh., et al. IT -ecosystem of the HybriLIT heterogeneous platform for high \nperformance comp uting and training of IT -specialists // CEUR Workshop Proceedings. \n2018 . 2267 , 638-644. \n9. Bashashin M.V., Zemlyanaya E.V. Comparative performance analysis of MPI - and \nOpenMP -programs on the example of parallel calculations in the framework of the nucleus -\nnucleus potential model and the φ0-spintronic model // Modern Information Technologies \nand IT -Education. 2022. Vol. 18, №3 , 545-557. \n " }, { "title": "1906.00249v1.Magnetic_fluctuations_in_the_itinerant_ferromagnet_LaCrGe3_studied_by_139La_NMR.pdf", "content": "arXiv:1906.00249v1 [cond-mat.str-el] 1 Jun 2019Magnetic fluctuations in the itinerant ferromagnet LaCrGe 3studied by139La NMR\nK. Rana,1H. Kotegawa,2R. R. Ullah,3J. S. Harvey,3S. L.\nBud’ko,1P. C. Canfield,1H. Tou,2V. Taufour,3and Y. Furukawa1\n1Ames Laboratory, U.S. DOE, and Department of Physics and Ast ronomy, Iowa State University, Ames, Iowa 50011, USA\n2Graduate School of Science, Kobe University, Kobe 657-8501 , Japan\n3Department of Physics, University of California, Davis, CA 95616, USA\n(Dated: June 4, 2019)\nLaCrGe 3is an itinerant ferromagnet with a Curie temperature of Tc= 85 K and exhibits an\navoided ferromagnetic quantum critical point under pressu re through a modulated antiferromag-\nnetic phase as well as tri-critical wing structure in its tem perature-pressure-magnetic field ( T-p-H)\nphase diagram. In order to understand the static and dynamic al magnetic properties of LaCrGe 3,\nwe carried out139La nuclear magnetic resonance (NMR) measurements. Based on the analysis\nof NMR data, using the self-consistent-renomalization (SC R) theory, the spin fluctuations in the\nparamagnetic state are revealed to be isotropic ferromagne tic and three dimensional (3D) in nature.\nMoreover, the system is found to follow the generalized Rhod es-Wohfarth relation which is expected\nin 3D itinerant ferromagnetic systems. As compared to other similar itinerant ferromagnets, the\nCr 3delectrons and their spin fluctuations are characterized to h ave a relatively high degree of\nlocalization in real space.\nPACS numbers:\nI. INTRODUCTION\nRecently muchattention has been paid toitinerant fer-\nromagnetic(FM) compounds because ofthe observations\nofunconventionalsuperconductivity(SC) aswellaschar-\nacteristic magnetic properties related to FM quantum\ncriticality under application of pressure ( p) and magnetic\nfield (H)[1–7]. Interestingly, in itinerant FM compounds,\nthe FM quantum critical point (QCP) under pis always\navoided, which has been of great interest in experimen-\ntal and theoretical studies. Usually, when the second\norder paramagnetic (PM)-FM phase transition temper-\nature (Tc) is suppressed by the application of p, the or-\nder of the phase transition changes to the first order at\nthe trictrtical point (TCP) before Tcreaches 0 K at the\nquantum phase transition (QPT). This is known as the\navoided QCP [8–10]. Here the QCP is a second order\nquantum phase transition at T= 0 K. [11] When the\nPM-FM transition becomes of the first order at the TCP\nin thep-Tplane, the application of magnetic field ( H)\nleads to a tricritical wing (TCW) structure in the T-p-H\nthree dimensional phasediagram[see, Fig. 1(a)] asfound\nin UGe 2[8, 9] and ZrZn 2[10]. A PM-FM QCP can also\nbe avoided by the appearance of an antiferromagnetic\n(AFM) ordered state under pnear the putative QCP, as\nactually observed in CeRuPO[12, 13] and MnP[14, 15].\nIn this case, no wing structure has been reported and the\nAFM state is suppressed by the application of moderate\nH, as schematically shown in Fig. 1(b).\nSuch interesting phase diagrams have been theoreti-\ncally well studied, and PM-FM QCP in clean itinerant\nFM systems is suggested to be always avoided following\nthe above two scenarios. It has also been suggested that\nthe coupling of quantum fluctuations of particle-hole ex-\ncitations with the order parameter in metallic systems\ncan generically render the prior second order phase tran-\nFIG. 1: Generic temperature-pressure-magnetic field ( T-p-\nH) phase diagrams for clean itinerant ferromagnets. (a)\nSchematic T-p-Hphase diagram with a tricritical wing\n(TCW) structure. The second order phase transition (solid\nline) becomes first order (dashed line) at a tricritical poin t\n(TCP). Under magnetic field, the TCW structure shown in\nred emerge from the TCP. (b) Schematic T-p-Hphase dia-\ngram with an antiferromagnetic (AFM) phase in blue. The\nAFM phase emerges from the Lifshitz point (LP) that is sup-\npressed under H.\nsitions first order [16–20] or drive the system towards\nincommensurate ordering states [17, 18]. In addition,\nmagnetic fluctuations may also play an important role\nto avoid FM QCP as some systems exhibit phases such\nas incommensurate magnetic ordered states [12–15] and\nunconventional superconductivity [3–7]. Therefore, the\nexperimental characterization of the nature of quantum\nfluctuations, including magnetic fluctuations, in these2\nclasses of materials is important to illuminate the un-\nderlying mechanism that connects magnetism, quantum\ncritically and superconductivity.\nRecently, the itinerant ferromagnet LaCrGe 3has been\ndiscovered to be a new class of itinerant ferromagnets\nexhibiting the remarkable T-p-Hphase diagram where\nboth the TCW structure and AFM phase are observed\n[21–23]. LaCrGe 3crystallizes in the hexagonal BaNiO 3-\ntype structure [space group P63/mmc(194)][24]. At am-\nbientp, LaCrGe 3is FM below the Curie temperature Tc\n= 85 K with an ordered magnetic moment at low tem-\nperatures of 1.25 µB/Cr aligned along the caxis [21].\nThis small value of the magnetic moment compared with\nthe effective moment above Tc(µeff= 2.4µB/Cr) [25] in\nthe PM state suggests some degree of delocalization of\nthe Cr 3dspins.Tccan be suppressed with pleading to\na weakly modulated AFM state around 1.5 GPa and T\nclose to 50 K [21]. Furthermore, it was also reported to\nexhibit a TCP where the second order FM transition be-\ncomes of the first order at Tof around 40 K and pclose\nto 1.8 GPa [22] yielding a TCW structure under H.\nMotivated by the novel magnetic properties in\nLaCrGe 3, we carried out nuclear magnetic resonance\n(NMR) measurement which is a powerful technique to\ninvestigate the magnetic and electronic properties of ma-\nterials from a microscopic point of view. It is known that\nthe temperature dependence of the nuclear spin-lattice\nrelaxation rate (1/ T1) reflects the wave vector q-summed\ndynamical susceptibility. On the other hand, NMR spec-\ntrum measurements, in particular the Knight shift K,\ngive us information on local static magnetic susceptibil-\nityχ. Thus from the temperature dependence of 1/ T1T\nandK, onecan obtain valuableinsights intospin fluctua-\ntions in materials. In this paper, we report the results of\n139La NMR measurements performed to investigate the\nspin fluctuations in LaCrGe 3. Our analysis, based on the\nself-consistent renomalization (SCR) theory, reveals that\nFM spin fluctuations due to Cr 3 dspins are of3D nature.\nIn addition, LaCrGe 3is well characterizedby a relatively\nhigh degree of localization of 3 dspins although the sys-\ntem is itinerant. Furthermore, the spin fluctuations are\nalsorevealedtohaveamorelocalizednaturein realspace\ncompared to other similar itinerant FM materials.\nII. EXPERIMENTAL DETAILS\nNeedle-like shaped single crystals of LaCrGe 3were\ngrown out of high temperature solutions, the details of\nwhich are reported in Ref. [25]. Plural single crystals\nused for NMR measurements were placed in parallel on\na glass plate (5 ×5×0.2 mm3) to align their directions.\nThe crystalline caxis and the abplane are parallel and\nperpendicular to the needle direction of the crystal, re-\nspectively. An NMR coil was tightly wound around the\ncrystals including the glass plate to reduce a loss of the\nfilling factor that is estimated to be about 0.6. NMR\nmeasurements of139La (I=7\n2,γN\n2π= 6.0146 MHz/T,Q= 0.21 barns) nuclei were conducted using a lab-built\nphase-coherent spin-echo pulse spectrometer. The139La\nNMR spectra were obtained by sweeping Hat fixed fre-\nquencies or by sweeping frequency under constant H.H\nwas applied parallel to either the crystalline caxis or\ntheabplane. The zero-shift position corresponding to\nthe Larmor field for each resonance frequency was deter-\nmined by31P NMR in H 3PO4solution or63Cu NMR in\nCu metal.\nThe139La nuclear spin-lattice relaxation rate (1/ T1)\nwas measured with a saturation recovery method. 1 /T1\nat each temperature ( T) was determined by fitting the\nnuclear magnetization Mversus time tusing the ex-\nponential function 1 −M(t)/M(∞) = 0.012e−t/T1+\n0.068e−6t/T1+0.206e−15t/T1+0.714e−28t/T1, whereM(t)\nandM(∞) are the nuclear magnetization at time tafter\nthesaturationandtheequilibrium nuclearmagnetization\natt→ ∞, respectively, for the case of magnetic relax-\nation [26]. The observed recovery data in the paramag-\nnetic state were well fitted by the function, indicating\nthat the nuclear relaxation is mainly induced by fluctu-\nations of the hyperfine field at the139La site. For the\nanalysis of NMR data, we measured the magnetic sus-\nceptibility χ(T) of the single crystal at H= 7 T applied\nparallel to the caxis and to the abplane in a commercial\nQuantum Design superconducting quantum interference\ndevice magnetometer.\nIII. RESULTS AND DISCUSSION\nA.139La NMR spectrum\nFigure 2(a) shows the field-swept139La-NMR spectra\nof LaCrGe 3atT= 230 K for Hparallel to the caxis\n(H||c) and to the abplane (H||ab). The typical NMR\nspectrum for a nucleus with spin I= 7/2 with Zeeman\nand quadrupolar interactions can be described by a nu-\nclear spin Hamiltonian H=−γ¯hI·Heff+hνQ\n6[3I2\nz−I2+\n1\n2η(I2\n++I2\n−)], where Heffis the effective field at the nu-\nclear site, his Planck’s constant, and ηis the asymmetry\nparameter of electric field gradient (EFG) at the nuclear\nsite. The nuclear quadrupole frequency for I= 7/2 nu-\nclei is given by νQ=e2QVZZ/14h, whereQis the nu-\nclear quadrupole moment and VZZis the EFG at the La\nsite. When the Zeeman interaction is much greater than\nthe Quadrupole interaction, this Hamiltonian produces a\nspectrum with a central transition line flanked by three\nsatellite peaks on either side. The observed spectra are\nwell reproduced by simulated spectra (red lines) from the\nsimple Hamiltonian with νQ= 0.66 MHz and η∼0. The\ntiny extra peaks around 6.80 T and 7.66 T for H||cin\nFig. 1(a) could be due to mis-orientation of some of the\ncrystals while being attached on the glass plate and also\nmay be due to slightly different qualities of the crystals.\nThe values of νQandηestimated from the main seven\npeaks are found to be independent of temperature in the\nPM state above TC= 85 K. From the spectrum analy-3\nFIG. 2: (a) Typical field-swept139La-NMR spectra of\nLaCrGe 3in the PM state at f= 42.5 MHz and T= 230\nK forH||candH||ab. (b) Temperature dependence of139La-\nNMR spectrum for H||c. (c) Field-swept139La-NMR spectra\nin the FM state ( T= 1.65 K) for H||candH||ab. Blue and\ngreen curves represent spectra for H||candH||ab, respec-\ntively, and the red curves are the calculated spectra with νQ\n= 0.66 MHz for the PM state and νQ∼0.49 MHz for the FM\nstate. The vertical dashed black lines in (a) and (b) represe nt\nthe zero-shift position ( K= 0).\nsis where θs are found to be 0 and π/2 forH/bardblcand\nH/bardblab, respectively, it is clear that the principal axis\nof the EFG at the La site is along the caxis. Since the\nLa site in LaGeCr 3does not have a local axial symmetry\n(¯6m2), one may expect finite value of η. However, ηis\nfound to be very close to zero within our experimental\nuncertainty.\nAs shown in Fig. 2(b), with decreasing temperatures,\neach line becomes broader due to inhomogeneous mag-\nnetic broadening and the spectra show less clear features\nof the quadrupolar split lines below ∼155 K. At the\nsame time, nuclear spin-spin relaxation time T2becomes\nshort at a wide range of temperatures close to TC. Those\nmakeNMRspectrummeasurementsdifficultbelow ∼110\nK. However, when the temperature is decreased down to\n1.65K,wellbelow TC=85K,wewereabletoobservethe\n139La NMR spectrum in the FM state as shown at the\nbottom of Fig. 2(b), where the spectrum largely shifted\nto higher magnetic field by ∼4 T. The shift is due to\nthe internal magnetic induction ( Bint) at the La site pro-\nduced by the Cr spontaneous magnetic moments in the\nFM state. The νQ∼0.49 MHz estimated from the spec-\ntra under two different magnetic field directions ( H||c\nandH||ab) shown in Fig. 2(c) is slightly smaller than\n0.66 MHz observed in the PM state. From the spectrum,\ntheBintis estimated to be −4 T and−4.2 T for H||cand\nH||ab, respectively. Here we took the zero-shift position\n(K= 0) as the origin of Bint. Unfortunately, the spec-\ntrum is measurable only around 1.6 K since its intensity/s45/s51/s46/s48 /s45/s51/s46/s53 /s45/s52/s46/s48 /s45/s52/s46/s53/s72 /s124/s124 /s99\n/s32/s61/s32/s48/s111\n/s102/s32/s61/s32/s49/s56/s46/s56/s49/s32/s77/s72/s122\n/s102/s32/s61/s32/s50/s50/s46/s55/s32/s77/s72/s122\n/s102/s32 /s61/s32/s51/s49/s32/s77/s72/s122\n/s32/s32/s83/s112/s105/s110/s32/s101/s99/s104/s111/s32/s105/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s114/s98/s46/s32/s117/s110/s105/s116/s115/s41/s32\n/s102/s47\n/s78/s45/s72 /s32/s40/s84/s41/s102/s32/s61/s32/s52/s50/s46/s53/s32/s77/s72/s122/s54 /s56 /s49/s48 /s49/s50/s45/s51/s46/s48/s45/s51/s46/s53/s45/s52/s46/s48/s45/s52/s46/s53/s45/s53/s46/s48\n/s32/s72 /s124/s124 /s99\n/s32/s72 /s124/s124 /s97/s98\n/s32/s32/s66\n/s105/s110/s116/s32/s40/s84/s41\n/s72\n/s114/s101/s115/s32/s32/s40/s84/s41\nFIG. 3: (a)139LaNMR spectra at various frequencies for H||c\natT= 1.65 K. Inset: Resonance magnetic field ( Hres) depen-\ndence of BintforH||ab(green circles) and H||c(blue squares).\nThe arrows show the positions of the central transition line\nof the spectra measured at different frequencies\ndecreases rapidly by raising the temperature.\nIn order to check whether the Bintis induced by the\nspontaneous Cr magnetic moments in the FM state and\nnot due to NMR shift produced by the application of\nmagnetic field, wedetermined the externalmagnetic field\ndependence of Bintby measuring the spectra with differ-\nent frequencies as shown in Fig. 3. Although the signal\nintensity decreases with decreasing resonance frequency,\nwe observed the spectrum down to f= 18.81 MHz and\nfound that Bint∼ −4 T and −4.2 T for H||candH||ab\nare nearly independent of resonance frequency (i.e. ex-\nternal magnetic field), confirming that these Bintvalues\ncan be attributed to the hyperfine field at the La sites\nproduced by the magnetic field independent spontaneous\nmagnetic moment of the Cr ions in LaCrGe 3in the FM\nstate.\nFigure 4 shows the Tdependence of the139La-NMR\nshift in the PM state for H/bardblabplane (Kab) andH/bardblc\naxis (Kc) determined from the simulated spectra, where\nbothKabandKcare nearly the same and decrease on\nlowering temperature. The NMR shift consists of tem-\nperature dependent spin shift Ks(T) andTindependent\norbital shift K0:K(T) =Ks(T) +K0whereKs(T) is4\n/s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s45/s51/s48/s45/s50/s53/s45/s50/s48/s45/s49/s53/s45/s49/s48/s45/s53/s48\n/s84 /s32/s32/s40/s75/s41/s32\n/s32/s75\n/s99/s44/s32 /s75\n/s97/s98/s32/s40/s37/s41/s32/s75\n/s97/s98\n/s32/s75\n/s99\n/s48/s46/s48/s48 /s48/s46/s48/s49 /s48/s46/s48/s50 /s48/s46/s48/s51 /s48/s46/s48/s52/s45/s50/s53/s45/s50/s48/s45/s49/s53/s45/s49/s48/s45/s53/s48\n/s32/s32/s75\n/s99/s44/s32 /s75\n/s97/s98/s32/s40/s37/s41\n/s99/s44/s32\n/s97/s98/s32/s32/s40/s101/s109/s117/s47/s109/s111/s108/s41\nFIG. 4: Temperature dependence of139La Knight shift for\nbothH||ab(green) and H||c(blue) directions measured at H\n∼7 T. Inset: Kabvs.χabandKcvs.χcplots. The solid\nlines are linear fits as described in the text.\nproportional to the spin part of magnetic susceptibil-\nityχs(T) via hyperfine coupling constant A,Ks(T) =\nAχs(T)\nNA. HereNAis Avogadro’s number. The hyperfine\ncoupling constants are estimated to be Aab=−32±1\nkOe/µBandAc=−27±1 kOe/µBforH||abandH||c,\nrespectively, from the slopes in the so-called K-χplots\nshown in the inset of Figure 4. The intercepts for the fits\nare almost zero for both the directions. This indicates\nthat the observed Knight shifts are mainly attributed to\nKs(T).Bintis proportional to Ahf<µ>whereAhfis the\nhyperfine coupling constant and <µ>is the ordered Cr\nmagnetic moment. Using Bint=−4.0 T and −4.2 T for\nH||candH||ab, respectively, and Aab=−32kOe/µBand\nAc=−27 kOe/µB,<µ>are estimated to be 1.30 µBfor\nH||aband 1.50 µBforH||cwhich are slightly higher but\nin good agreement with 1.22 µB(along the caxis) re-\nported by the neutron diffraction measurements [27] and\n1.25µB(along the caxis) from magnetization measure-\nments [23, 25].\nB.139La spin lattice relaxation time\nIn order to investigate the magnetic fluctuations in\nLaCrGe 3, we measured the139La spin-lattice relaxation\nrate (1/T1) at the peak position of the spectra for both\nthe magnetic field directions. Figure 5 showsthe temper-\nature dependence of 1/ T1Twhere 1/ T1Tincreases with\ndecreasing temperature from room temperature to 125 K\nwith no anisotropy in T1.\nBased on the T1and the spin part of the Knight shift\n(Ks) data, we discuss the magnetic fluctuations in the\nPM state of LaCrGe 3. First we tentatively employ the\nmodified Korringaratioanalysis. In Fermi liquid picture,\n1/T1TandKsare determined by the density of states at/s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s49/s48/s49/s48/s48/s49/s48/s48/s48\n/s49/s48/s49/s48/s48/s49/s48/s48/s48\n/s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s50/s52/s54/s56/s49/s48/s49/s50\n/s49/s47/s40 /s84\n/s49/s84/s75 /s41/s32/s32/s32/s32 /s105/s110/s45/s112/s108/s97/s110/s101/s44/s32 /s32/s111/s117/s116/s45/s111/s102/s45/s112/s108/s97/s110/s101\n/s49/s47/s40 /s84\n/s49/s84/s75/s51/s47/s50\n/s41/s32 /s32/s105/s110/s45/s112/s108/s97/s110/s101/s44/s32 /s32/s111/s117/s116/s45/s111/s102/s45/s112/s108/s97/s110/s101/s32\n/s32/s49/s47/s40 /s84\n/s49/s84/s75/s51/s47/s50\n/s41/s32/s32/s32/s40/s115/s75/s41/s45/s49/s49/s47/s40 /s84\n/s49/s84/s75 /s41/s32/s32/s32/s32/s40/s115/s75/s41/s45/s49\n/s84 /s32/s32/s40/s75/s41/s32/s49/s47 /s84\n/s49/s84 /s32/s32/s32/s32/s40/s115/s75/s41/s45/s49\n/s32/s72 /s124/s124 /s97/s98\n/s32/s72 /s124/s124 /s99\n/s84 /s32/s32/s40/s75/s41\nFIG. 5: Temperature dependence of139La 1/T1TforH||ab\n(green circles) and H||c(blue squares). The inset shows the\nsemi- log plots of 1/( T1TKs) and 1/( T1TK3/2\ns) for two differ-\nent magnetic fluctuation directions (in-plane and out-of-p lane\ndirections).\nthe Fermi energy D(EF). TheT1has relation with Ks\nthat can be described as T1TK2\ns= (¯h/4πkB)(γe/γn)2≡\nS. Hereγeis the electronic gyromagnetic ratio. The\nKorringa ratio α(≡S/T1TK2\ns) between an experimen-\ntal value of T1TK2\nsand the non-interacting electron sys-\ntemScan reveal information about electron correlations\nin materials [28, 29]. α∼1 represents the situation\nof uncorrelated electrons. However, enhancement of χs\n(q/negationslash= 0) increases 1/ T1Tbut has little or no effect on\nKswhich probes only the uniform χs(q= 0) where q\nrepresents wave vector. Thus α >1 indicates antifer-\nromagnetic (AFM) spin correlations. In contrast, FM\nspin correlations produce α <1. Since 1/ T1Tprobes\nmagneticfluctuationsperpendiculartothe magneticfield\n[28], in general, one should consider the Korringa ratio\n1/(T1,⊥TK2\ns,ab), where 1/( T1,⊥T) = 1/(T1T)H||c, when\nexamining the character of magnetic fluctuations in the\nabplane (in-plane direction). One also needs to con-\nsider the Korringa ratio 1/( T1,||TK2\ns,c) for magnetic fluc-\ntuations along the caxis (out-of-plane direction). Here,\n1/(T1,/bardblT) is estimated from 2/( T1T)H||ab−1/(T1T)H||c.\n[30]. However, since both 1/ T1and139Kare nearly\nisotropic, 1/( T1,⊥TK2\ns,ab) and 1/( T1,||TK2\ns,c) are almost\nthe same, clearly indicating isotropic magnetic fluctua-\ntions in LaCrGe 3.α/bardblandα⊥decrease from ∼0.07 at\nroom temperature to less than ∼0.008 around 120 K, in-\ndicatingdominantFM spincorrelationsbetweenCrspins\nin the compound.\nItshouldbenotedthat, however,theKorringaanalysis\nusually applies for PM materials where electron-electron\ninteraction is weak. Since LaCrGe 3exhibits a FM or-\nder, we also analyze NMR data based on self-consistent\nrenormalization(SCR)theory. Asshownabove,themag-\nnetic fluctuations are governed by FM spin correlations.5\nIn this case, according to SCR theory for weak itiner-\nant ferromagnets, 1/( T1TKs) and 1/( T1TK3/2\ns) are ex-\npected to be independent of Tfor three dimensional (3D)\nor two-dimensional (2D) FM spin fluctuations, respec-\ntively [31, 32]. The inset of Fig. 5 shows the Tde-\npendence of 1/( T1TKs) and 1/( T1TK3/2\ns) for the two\ndirections. Both the 1/( T1,/bardblTKs,c) and 1/( T1,⊥TKs,ab)\nare nearly constant, while both the 1/( T1,/bardblTK3/2\ns,c) and\n1/(T1,⊥TK3/2\ns,ab) increase with increasing temperature.\nThis indicates that the FM spin fluctuations are char-\nacterized as 3D in nature.\nC. Spin fluctuations\nThe Curie Weiss behavior of 1/ T1Tin itinerant ferro-\nmagents is well described in the SCR theory in terms of\n3DFM spin fluctuations [33–37], as described above. Ac-\ncording to SCR theory, there are two important param-\neters to characterize spin fluctuation for itinerant ferro-\nmagnets: T0andTAcorresponding to the widths of the\nspin fluctuations in frequency ( ω) space at wave vector\n(q) = 0 and the width of the distribution of static suscep-\ntibility in qspace at ω= 0, respectively [38]. The former\ncan be obtained from the following equation of 1/ T1Tas\nderived by Corti et al.[39],\n1\nT1T=3¯hγ2\nNHhf\n16πµBK\nT0(1)\nHereHhfis the hyperfine field experienced at the La site\nper spin [39]. From the experimentally determined val-\nues,T0is estimated to be 89 K and 75 K for the in-plane\nand out-of-plane directions of spin fluctuations, respec-\ntively. On the other hand, TAcan be estimated using the\nfollowing equation given by Takahashi [38]:\nTA= 20C4/3/bracketleftBigTc\np3/2\nsT1/4\n0/bracketrightBig4/3\n(2)\nwhereTc= 85 K, psis the saturated moment and\nC4/3=1.006089 is a constant. Utilizing these values, TA\nis estimated to be 998 K and 793 K for the in-plane and\nout-of-plane directions of spin fluctuations, respectively.\nThe estimated values of TAseem to be comparable or\nslightly greater than those in uranium based compounds\nsuch as URhGe ( TA= 568 K) [40] and UGe 2(TA= 442\nK) [40], but are smaller than those in 3 delectron itin-\nerant ferromagnets such as ZrZn 2(TA= 7400 K) [38],\nNiAl3(TA= 3670 K) [38] and MnSi ( TA= 1690 K) [39].\nIn order to compare LaCrGe 3with other clean itiner-\nant ferromagnets that avoid the FM QCP with the TCW\nstructure, we plotted their spin fluctuation parametersin\nFig. 6 along with those in LaCrGe 3. The plot is known\nas the Generalized Rhodes-Wohlfarth plot. The xaxis\nof this plot is the ratio of TCandT0and theyaxis the\nratio of the effective paramagnetic moment ( peff) andps./s48/s46/s48 /s48/s46/s53 /s49/s46/s48 /s49/s46/s53/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52\n/s76/s97/s67/s114/s71/s101\n/s51/s32/s79/s80/s85/s82/s104/s71/s101\n/s32/s32/s77/s110/s83/s105\n/s32/s90/s114/s90/s110\n/s50/s32/s32\n/s85/s82/s104/s65/s108/s32 /s32\n/s32/s112\n/s101/s102/s102/s32/s47/s32/s112\n/s115\n/s84\n/s99/s32/s47/s32 /s84\n/s48/s85/s71/s101\n/s50/s32\n/s76/s97/s67/s114/s71/s101\n/s51/s32/s73/s80\nFIG. 6: Generalized Rhodes-Wohlfarth plot for the in-plane\n(blue) and out-of-plane (green) directions for LaCrGe 3along\nwith other itinerant ferromagnets that avoid FM-PM QCP\nthrough tricritical wings under pressure. The values of pa-\nrameters are estimated from Ref. [38] for ZrZn 2and MnSi\nand Ref. [40] for UGe 2, URhGe and URhAl. The red line\nrepresents the generalized Rhodes-Wohlfarth relation in t he\n3D system: peff/ps= 1.4 (TC/T0)−2/3[38].\nAccording to SCR theory, in the case of itinerant ferro-\nmagnets, magnetic fluctuations contribute to peffin PM\nstate. Therefore, one can expect that the ratio of peff/ps\nbecomes large, when spin fluctuations are important in\nthese systems. As for T0, based on SCR theory, TC/T0is\nclose to unity in the case of localized ferromagnets, while\nthe ratio becomes less than unity if itinerant character\nbecomes significant. These situations are predicted the-\noretically through the generalized Rhodes-Wolfarth rela-\ntion ofpeff/ps= 1.4 (TC/T0)−2/3(red line), and the rela-\ntion has shown great empirical agreement with materials\nhaving 3D FM fluctuations [38]. As shown in the Fig. 6,\nLaCrGe 3and other typical itinerant ferromagnets seem\nto follow the relation. However, the values of LaCrGe 3\nfor both the in-plane and out-of-plane directions of mag-\nnetic fluctuations are located close to unity, indicating a\nrelatively high degree of localization in Cr 3 delectrons\neven though the system is itinerant. In addition, the re-\nsults also indicate that the spin fluctuations in LaCrGe 3\nshow a more localized nature in real space than the other\nsimilar itinerant FM materials compared.\nIV. SUMMARY\nIn summary, we carried out139La NMR in the itiner-\nant ferromagnet LaCrGe 3to characterize the magnetic\nproperties from a microscopic point of view. The prin-\ncipal axis of the electric field gradient at the La site has\nbeen shown to be alongthe caxis with νQof 0.66 MHz in\nthe PMstate andabout 0.49MHz in the FM state.139La6\nNMR spectra measurements in the FM state confirmed\nthe FM ordered state below TC= 85 K with a magnetic\nmomentof ∼1.4µB/Cr,consistentwithpreviousreports.\nBased on the Korringaratio analysisand the SCR theory\nusing the resultsof 1/ T1TandK, spin fluctuations in the\nPM state are revealed to be isotropic FM and three di-\nmensional in nature. In addition, the system is found to\nfollow the generalized Rhodes-Wohfarth relation which\nis expected in 3D itinerant FM systems. TC/T0is found\nto be close to unity, indicating that there is a relatively\nhigh degree of localization in the 3 dCr electrons even\nthough the system is itinerant. It would be interesting\nif these uniqueness in magnetic fluctuations in LaCrGe 3\nis related to the appearance of both tricritcal wings and\nAFM state under pressure and magnetic fields. Analysis\nof spin fluctuations in those materials where FM QCP is\navoided by the appearance of AFM phase, and further\nstudies on LaCrGe 3under pressure are important to in-\nvestigate this connection, which are now in progress.V. ACKNOWLEDGMENTS\nThe authors would like to thank Q.-P. Ding, R.\nTakeuchi and Y. Kuwata, for help in conducting exper-\niments and T. Matsui for fruitful discussions. The re-\nsearch was supported by the U.S. Department of Energy,\nOffice of Basic Energy Sciences, Division of Materials\nSciences and Engineering. Ames Laboratory is oper-\nated for the U.S. Department of Energy by Iowa State\nUniversity under Contract No. DE-AC02-07CH11358.\nPart of the work was supported by the Japan So-\nciety for the Promotion of Science KAKENHI Grant\nNumbers JP15H05882, JP15H05885, JP15K21732, and\nJP18H04321 (J-Physics). K. R. also thanks the KAK-\nENHI: J-Physics for the financial support that provided\nan opportunity to be a visiting scholar at Kobe Univer-\nsity.\n[1] P. C. Canfield, and S. L. Bud’ko, Preserved entropy and\nfragile magnetism, Rep. Prog. Phys. 79, 084506 (2016).\n[2] M. Brando, D. Belitz, F.M. Grosche, and T. R. Kirk-\npatrick, Metallic quantum ferromagnets, Rev. Mod.\nPhys.88, 025006 (2016).\n[3] E. A. Yelland, J.M. Barraclough, W. Wang, K.V.\nKamenev, and A.D. Huxley, High-field superconductiv-\nity at an electronic topological transition in URhGe, Nat.\nPhys.7, 890 (2011).\n[4] S.S.Saxena, P.Agarwal, K.Ahilan, F.M. Grosche, R.K.\nW. Haselwimmer, M. J. Steiner, E. Pugh, I. R. Walker,\nS. R. Julian, P. Monthoux, G. G. Lonzarich, A. Huxley, I.\nSheikin, D. Braithwaite, and J. Flouquet, Superconduc-\ntivity on the border of itinerant-electron ferromagnetism\nin UGe 2, Nature (London), 406, 587 (2000).\n[5] E. Hassinger, D. Aoki, G. Knebel, and J. Flouquet,\nPressuretemperature Phase Diagram of Polycrystalline\nUCoGe Studied by Resistivity Measurement, J. Phys.\nSoc. Jpn., 77, 073703 (2008).\n[6] D. Aoki, A.D. Huxley, E. Ressouche, D. Braithwaite,\nJ. Flouquet, J. Brison, E. Lhotel, and C. Paulen, Co-\nexistence of superconductivity and ferromagnetism in\nURhGe, Nature (London) 413, 613 (2001).\n[7] N. T. Huy, A. Gasparini, D..E. de Nijs, Y. Huang, J.C.P.\nKlaasse, T. Gortenmulder, A. de Visser, A. Hamann,\nT. G¨ orlach, and H. v. L¨ ohneysen, Superconductivity on\nthe Border of Weak Itinerant Ferromagnetism in UCoGe,\nPhys. Rev. Lett. 99, 067006 (2007).\n[8] V. Taufour, D. Aoki, G. Knebel, and J. Flouquet, Tri-\ncritical Point and Wing Structure in the Itinerant Ferro-\nmagnet UGe 2, Phys. Rev. Lett. 105, 217201 (2010).\n[9] H. Kotegawa, V. Taufour, D. Aoki, G. Knebel, and J.\nFlouquet, Evolution toward Quantum Critical end Point\nin UGe 2, J. Phys. Soc. Jpn. 80, 083703 (2011).\n[10] N. Kabeya, H. Maekawa, K. Deguchi, N. Kimura, H.\nAoki, and N. K. Sato, Non-Fermi Liquid State Bounded\nby a Possible Electronic Topological Transition in ZrZn 2,\nJ. Phys. Soc. Jpn. 81, 073706 (2012).\n[11] S. Sachdev, Quantum Phase Transitions , (CambridgeUniversity Press, New York, 2011).\n[12] H. Kotegawa, T. Toyama, S. Kitagawa, H. Tou, R. Ya-\nmauchi, E. Matsuoka, and H. Sugawara, PressureTem-\nperatureMagnetic Field Phase Diagram of Ferromagnetic\nKondo Lattice CeRuPO, J. Phys. Soc. Jpn. 82, 123711\n(2013).\n[13] E. Lengyel, M. E. Macovei, A. Jesche, C. Krellner, C.\nGeibel, and M. Nicklas, Avoided ferromagnetic quan-\ntum critical point in CeRuPO, Phys. Rev. B 91, 035130\n(2015).\n[14] J.G. Cheng, K. Matsubayashi, W. Wu, J.P. Sun, F. K.\nLin, J. L. Luo, and Y. Uwatoko, Pressure Induced Super-\nconductivity on the border of Magnetic Order in MnP,\nPhys. Rev. Lett. 114, 117001 (2015).\n[15] M. Matsuda, F. Ye, S. E. Dissanayake, J.-G. Cheng, S.\nChi, J. Ma, H. D. Zhou, J.-Q. Yan, S. Kasamatsu, O.\nSugino, T. Kato, K. Matsubayashi, T. Okada, and Y.\nUwatoko, Pressure dependence of the magnetic ground\nstates in MnP, Phys. Rev. B 93, 100405(R) (2016).\n[16] D. Belitz, T.R. Kirkpatrick and T. Vojta, First Order\nTransitions and Multicritical Points in Weak Itinerant\nFerromagnets, Phys. Rev. Lett. 82, 4707 (1999).\n[17] A. V. Chubukov and C. P´ epin, Instability of the\nQuantum-Critical Point of Itinerant Ferromagnets, J.\nRech, Phys. Rev. Lett. 92, 147003 (2004).\n[18] T. R. Kirkpatrick, and D. Belitz, Universal low-\ntemperaturetricritical pointinmetallic ferromagnets an d\nferrimagnets, Phys. Rev. B 85, 134451 (2012).\n[19] F. Kr¨ uger, C. J. Pedder, and A.G. Green, Fluctuation-\nDriven Magnetic Hard-Axis Ordering in Metallic Ferro-\nmagnets, Phys. Rev. Lett. 113, 147001 (2014).\n[20] Y. Sang, D. Belitz, and T. R. Kirkpatrick, Disorder De-\npendence of the Ferromagnetic Quantum Phase Transi-\ntion, Phys. Rev. Lett. 113, 207201 (2014).\n[21] V. Taufour, U. S. Kaluarachchi, R. Khasanov, M. C.\nNguyen, Z. Guguchia, P. K. Biswas, P. Bonf` a, Roberto\nDe Renzi, X. Lin, S. K. Kim, E. D. Mun, H. Kim,\nY. Furukawa, C.-Z. Wang, K.M. Ho, S. L. Budko, and\nP. C. Canfield, Ferromagnetic Quantum Critical Point7\nAvoided by the Appearance of Another Magnetic Phase\ninLaCrGe 3underPressure, Phys.Rev.Lett. 117, 037207\n(2016).\n[22] U.S. Kaluarachchi, S. L. Bud’ko, P.C. Canfield, and\nV. Taufour, Tricritical wings and modulated magnetic\nphases in LaCrGe 3under pressure, Nat. Comm. 8, 546\n(2017).\n[23] V. Taufour, U. S. Kaluarachchi, S. L. Budko, and P.\nC. Canfield, Ferromagnetic quantum criticality: New as-\npects from the phase diagram of LaCrGe 3, Physica B\n536, 483 (2018).\n[24] H. Bie, O. Y. Zelinska, A. V. Tkachuk, and A.\nMar, Structures and Physical Properties of Rare-Earth\nChromium Germanides RECrGe 3(RE = La-Nd, Sm),\nChem. Mater. 19, 4613 (2007).\n[25] X.Lin, V.Taufour, S.L.Bud’ko, andP.C. Canfield, Sup-\npression offerromagnetism intheLaV xCr1−xGe3system,\nPhys. Rev. B 88, 094405 (2013).\n[26] A. Narath, Nuclear Spin-Lattice Relaxation in Hexago-\nnal Transition Metals: Titanium, Phys. Rev. 162, 320\n(1967).\n[27] J. M. Cadogan, P. Lemoine, B.R. Slater, A. Mar, and\nM. Avdeev, Neutron Diffraction Study of the Hexagonal\nPerovskite-Type Compound LaCrGe 3, Solid State Phe-\nnom.194, 71 (2013).\n[28] T. Moriya, The Effect of Electron-Electron Interaction\non the Nuclear Spin Relaxation in Metals, J. Phys. Soc.\nJpn.18, 516 (1963).\n[29] A.Narathand H.T. Weaver, Effects ofElectron-Electron\nInteractions on Nuclear Spin-Lattice Relaxation Rates\nand Knight Shifts in Alkali and Noble Metals, Phys. Rev.\n175, 373 (1968).\n[30] P. Wiecki, B. Roy, D. C. Johnston, S. L. Bud’ko, P.C.\nCanfield, and Y. Furukawa, Competing Magnetic Fluc-\ntuations in Iron Pnictide Superconductors: Role of Fer-\nromagnetic Spin Correlations Revealed by NMR, Phys.Rev. Lett. 115, 137001 (2015).\n[31] T. Moriya and K. Ueda, Nuclear magnetic relaxation in\nweakly ferro-and antiferromagnetic metals, Solid State\nCommun. 15, 169 (1974).\n[32] M. Hatatani and T. Moriya, Ferromagnetic Spin Fluc-\ntuations in Two-Dimensional Metals, J. Phys. Soc. Jpn.\n64, 3434 (1995).\n[33] T. Moriya and A. Kawabata, Effect of Spin Fluctuations\non Itinerant Electron Ferromagnetism, J. Phys. Soc. Jpn.\n34, 639 (1973).\n[34] M. Kontani, T. Hioki, and Y. Masuda, Nuclear magnetic\nrelaxation in itinerant electron ferromagnet ZrZn 2, Solid\nState Commun. 18, 1251 (1976).\n[35] Y. Takahashi and T. Moriya, Quantitative Aspects of the\nTheory of Weak Itinerant Ferromagnetism, J. Phys. Soc.\nJpn.54,1592 (1985).\n[36] Y. Takahashi, On the Origin of the Curie-Weiss Law of\nthe Magnetic Susceptibility in Itinerant Electron Ferro-\nmagnetism, J. Phys. Soc. Jpn. 55, 3553 (1986).\n[37] K. Yoshimura, M. Takigawa, Y. Takahashi, H. Yasuoka,\nand Y. Nakmura, NMR Study of Weakly Itinerant Fer-\nromagnetic Y(Co 1−xAlx)2, J. Phys. Soc. Jpn. 56, 1138\n(1987) .\n[38] Y. Takahashi, Spin Fluctuation Theory of Itinerant Elec-\ntron Magnetism , (Springer-Verlag, Berlin, 2013).\n[39] M. Corti, F. Carbone, M. Filibian, T. Jarlborg, A. A.\nNugroho, and P. Carretta, Spin dynamics in a weakly\nitinerant magnet from29Si NMR in MnSi, Phys. Rev. B\n75, 115111 (2007).\n[40] N. Tateiwa, J. Posp´ ıˇ sil, Y. Haga, H. Sakai, T. D. Mat-\nsuda, and E. Yamamoto, Itinerant ferromagnetism in\nactinide 5f-electron systems: Phenomenological analysis\nwith spin fluctuation theory, Phys. Rev. B 96, 035125\n(2017)." }, { "title": "0808.0892v2.Electron_spin_resonance_in_Kondo_systems.pdf", "content": "arXiv:0808.0892v2 [cond-mat.str-el] 15 Oct 2008Electron spin resonance in Kondo systems\nElihu Abrahams1and Peter W¨ olfle2\n1Center for Materials Theory, Serin Physics Laboratory, Rut gers University, Piscataway, NJ 08854-8019\n2Institut f¨ ur Theorie der Kondensierten Materie,\nUniversit¨ at Karlsruhe, D-76128 Karlsruhe, Germany\n(Dated: October 31, 2018)\nWe calculate the dynamical spin response of Kondo impurity a nd Kondo lattice systems within\na semiphenomenological Fermi liquid description, at low te mperatures T < T K, the Kondo tem-\nperature, and low magnetic fields B≪kBTK/gµB. Fermi liquid parameters are determined by\ncomparison (i) with microscopic theory (numerical renorma lization group) for the impurity model\nand (ii) with experiment for the lattice model. We find in the i mpurity case that the true impurity\nspin resonance has a width of the order of TKand disappears altogether if the g-factors of impurity\nspin and conduction electron spin are equal. However, there is an impurity-induced resonance con-\ntribution at the conduction electron resonance. The latter is broadened by spin lattice relaxation\nand is usually unobservable. In contrast, for the Anderson l attice in the Kondo regime we find\nasharpESR resonance line only slightly shifted from the local reso nance and broadened by spin\nlattice relaxation, the latter significantly reduced by bot h the effects of heavy fermion physics and\nferromagnetic fluctuations. We conjecture that our findings explain the sharp ESR-lines recently\nobserved in several heavy fermion compounds.\nPACS numbers: 71.27.+a, 75.20Hr, 76.30.-v\nI. INTRODUCTION\nThe Kondo effect is arguably the best-studied many-\nbody effect in condensed matter physics1. In its initial\nform2,3, it involves a local “impurity” spin in a d- orf-\norbital, antiferromagnetically coupled to the spins of a\nconduction band in a dilute magnetic alloy. At temper-\naturesTbelow the dynamically generated energy scale\nTK, the Kondo temperature, this interaction causes a lo-\ncal spin 1 /2 to be fully screened. This behavior should\nbe noticeable in the T-dependence of the spin dynam-\nics of the system, as probed by electron spin resonance\n(ESR). In fact, the local spin resonance in dilute Kondo\ncompounds at T≫TKhad been observed even before\nthe Kondo effect was understood. Afterward there were\na number of systematic experimental investigations and\nperturbative calculations for the ESR at T≫TKin di-\nlute Kondo systems4.\nAt low temperatures T≪TK, on the other hand, neu-\ntron scattering studies revealed the existence of a broad\nspin excitation peak of width TK, interpreted as the\nKondo bound state5. Within the isotropic s-dexchange\n(Kondo) model the total spin is conserved. Therefore, in\nthe limit of equal g-factors of local moments and conduc-\ntion electron spins one expects a single spin resonance\nline at all temperatures, only broadened by spin lattice\nrelaxation. As we shall show below, in this limit the\nweight of the broad local spin resonance tends to zero.\nIn several recent experiments6,7low-temperature ESR\nhas been observed in some heavy-fermion metals, in par-\nticular YRh 2Si2(YRS)8. The phase diagram of YRS has\na magnetic-field induced quantum critical point and is a\nmodel system for the study of quantum criticality in the\nKondolattice. Consequently, the observationofa narrow\nESR resonance in this compound aroused great interest,especially since it was commonly believed that heavy-\nfermion ESR would be unobservable due to an enormous\nintrinsic linewidth ∆ Bof order kBTK/gµB6. HereTK\nis the lattice coherence (“Kondo”) temperature for the\nonset of heavy-fermion behavior and gµBis the gyro-\nmagnetic ratio for the resonance. These were the first\nobservationsofESR in Kondolattice systems at T < T K.\nInYRS,theobservednarrowdysonian9ESRlineshape\nwas originally interpreted6as indicating that the reso-\nnance was due to local spins at the Yb sites. Therefore,\ninitially the authors speculated that the appearence of a\nnarrow ESR line might indicate the suppression of the\nKondo effect near the quantum critical point, since, as\nexplained above, carrying over Kondo impurity physics\nto the Kondo lattice one might expect the local spins to\nbe screened by the Kondo effect, giving rise only to a\nbroad spin excitation peak, too wide to be observed in\nESR experiments. However, a closer look10revealed that\nitinerant (heavy) electron ESR could give rise to a simi-\nlar line shape since the carrier diffusion in YRS is quite\nslow. Thus, whether the resonance was that of localized\nor itinerant spins remains an open question.\nNow, a common feature of the compounds in which\nESR has been observed appears to be the existence of\nferromagnetic fluctuations7These findings challenge our\nunderstanding of heavy fermion compounds: How does\na sharp electron spin resonance emerge despite Kondo\nscreeningand spin lattice relaxation, and why is this pro-\ncess influenced by ferromagnetic fluctuations? We shall\naddress these questions in the framework of Fermi liquid\ntheory, taking the relevant parameters from numerical\nstudies and experiment.2\nII. ANDERSON IMPURITY MODEL IN THE\nKONDO SCREENED REGIME\nIn the Kondo regime an impurity spin is screened by\nthe conduction electron spins at (or near) the impurity.\nThe dynamics of the impurity spin is governed by the\nenergy scale of the corresponding many-body resonance,\nthe Kondo temperature TK. Nonetheless the conduc-\ntion electrons in the vicinity of the impurity show the\ninfluence of the Kondo screened state in their dynami-\ncal behavior. In the Anderson model, the local spin is\nthat of a localized felectronWe assume that the Zeeman\nsplittings ωfandωcinduced by a magnetic field acting\non the local and conduction electron spins are small com-\npared to the Kondo temperature TK. Then the Kondo\nscreened state is only weakly perturbed by the magnetic\nfield. At temperatures T≪TK, the spin resonance be-\nhavior of the impurity may then be described by Fermi\nliquid theory11.\nWe start from the bare Anderson model Hamiltonian\nH=Hc+/summationdisplay\nk,σǫkσc+\nkσckσ+/summationdisplay\nσǫfσnfσ+Unf↑nf↓\n+V/summationdisplay\nk,σ(f+\nσckσ+h.c.), (1)\nwhereHcis the conduction electron Hamiltonian and\nc+\nkσ,f+\nσarecreationoperatorsofthe conductionelectrons\nin momentum and spin eigenstates ( kσ), and of electrons\nin the local flevel, respectively. The operator nfσ=\nf+\nσfσcounts the number of electrons on the local level,\nandǫfσ=ǫf−ωfσ/2, σ=±1.\nThe effect ofthe interaction Uis torenormalizethe pa-\nrameters ǫfσ,U,Vto/tildewideǫfσ,/tildewideU,/tildewideVintherenormalizedFermi\nliquid type low energy Hamiltonian, Eq. (1) with the\nrenormalized parameters that may be calculated using\nthenumericalrenormaliztiongroup(NRG)method12. To\nkeep the algebra simple, we assume particle-hole symme-\ntry in the following. Then /tildewideǫfσ=−(/tildewideU+ωfσ)/2. The\nhybridization of the local level with the conduction band\nleads to an f-level broadening /tildewideΓ =π/tildewideV2N0∼TKwith\nN0= 1/Wthe local conduction electron density of states\n(DOS) at the Fermi level (in the model with flat DOS, W\nis the bandwidth). The initially rather large bare level\nwidth is renormalized down to the very narrow width of\nthe Kondo resonance. The NRG calculation shows that\n/tildewideU=π/tildewideΓ and/tildewideV2is O(TK/TF), where TFis the Fermi\ntemperature of the conduction electrons.\nIn the framework of Fermi liquid theory, the interac-\ntion has two major consequences: (i) it gives rise to a\nmolecular field renormalizing the collective response of\nthe system (ii) it leads to a finite lifetime of quasiparti-\ncles. However, the quasiparticle relaxationrate is limited\nby the available phase space and vanishes quadratically\nas the excitation energy goes to zero. Therefore, at tem-\nperatures T≪TKthe Landau quasiparticles are well-\ndefined. The quasiparticle decay contributes to the spin\nrelaxation rate. As we shall show, the local moment re-laxation is governed by rapid spin flips on the frequency\nscaleofTK, occuringaspartofthemanybodyresonance.\nThen at temperatures T≪TKwe may neglect the addi-\ntional relaxation caused by the quasiparticle decay.\nWe now consider the effects of the molecular field\ncaused by the Fermi liquid interaction /tildewideU. We treat the\ninteraction term in the Hamiltonian in mean field ap-\nproximation: /tildewideUnf↑nf↓≈1\n2/tildewideU[∝an}bracketle{tnf∝an}bracketri}htnf−∝an}bracketle{tmf∝an}bracketri}htmf+const],\nwhere we defined the density and spin density operators\nnf=nf↑+nf↓andmf=nf↑−nf↓. In the case\nof particle-hole symmetry, when ∝an}bracketle{tnf∝an}bracketri}ht= 1, the density\nterm is cancelled by the single-particle energy. The spin-\ndensity term gives rise to an effective magnetic field so\nthat/tildewideǫfσ+1\n2/tildewideU[∝an}bracketle{tnf∝an}bracketri}ht−σ∝an}bracketle{tmf∝an}bracketri}ht] =−σωf, which amounts to\na doubling of the Zeeman energy. Here we have used\nthat the spin polarization is given by m=∝an}bracketle{tmf∝an}bracketri}ht=\nχ+−\nff(0)ωf/2 =ωf/π/tildewideΓ, withχ+−\nff(0) = 2/π/tildewideΓ the static\nsusceptibility of the local spin and /tildewideU=π/tildewideΓ, as obtained\nfrom NRG calculations. Then the local electron Green’s\nfunction, including the coupling to the conduction elec-\ntrons, is given by\nGfσ(iωn) = [iωn+σωf+i/tildewideΓsign(ωn)]−1(2)\nThe local spectral function is given by Afσ(ω) =\nImGfσ(ω+i0) =/tildewideΓ/[(ω+σωf)2+/tildewideΓ2], describing the\nKondo resonance. We see that in a magnetic field the\nresonance is shifted from its zero field position ω= 0 to\nthe spin dependent position ω=−σωf, which is double\nthe Zeeman shift.\nWe use the definitions ωf=gfµBB,ωc=gcµBBand\ntakeχ+−\nff, etc. to be response functions of spin 1 /2 oper-\nators. The dynamical conduction electron susceptibility\nχccis characterized by a resonance peak broadened by\nspin-lattice relaxation. We follow Barnes and Zitkova-\nWilcox13and model the spin-lattice relaxation mecha-\nnism by a local random magnetic field hithat fluctuates\nin both direction and magnitude. Then the conduction\nelectron Hamiltonian is\nHc=/summationdisplay\nk,σǫkσc+\nkσckσ\n+/summationdisplay\nk,σ,k′,σ′/summationdisplay\nihi·c+\nkσσσσ′ck′σ′ei(k−k′)·Ri(3)\nThe random field components are assumed to be Gaus-\nsian correlated, with ∝an}bracketle{thi∝an}bracketri}ht= 0 and ∝an}bracketle{thm\nihn\nj∝an}bracketri}ht=δijδmnh2.\nIn Born approximation the average conduction electron\nGreen’s function is then given by\nG0\nc↑(k,iωn) = [iωn−εk+ωc/2+iγsign(ωn)]−1,(4)\nwhereγ=πN0h2.\nThe impurity induced component of the dynamical\ntransverse susceptibility χ+−\nimp(Ω), where Ω is the fre-\nquency of an a.c. electromagnetic field polarized trans-\nverse to the static magnetic field, is a sum of three con-\ntributions, from the conduction electrons (cc), the local3\nelectrons (ff) and the mixed response of conduction elec-\ntrons to a spin polarization of the local electrons or vice\nversa (cf):\nχ+−\nimp(Ω) =µ2\nB{g2\nfχ+−\nff(Ω)+2gcgfχ+−\ncf(Ω)\n+g2\nc[χ+−\ncc(Ω)−χbulk\ncc(Ω)]} (5)\nThe partial susceptibilities may be calculated using stan-\ndardmanybodytechniques, seetheAppendix. Onefinds\nresonances at the two frequencies ωfandωc. They have\nvery different widths: the local electron spin resonance is\nbroadened by /tildewideΓ, whereas the bulk and the impurity in-\nduced conduction electron resonances are broadened by\n4γ. The results are given in the Appendix, Eqs. (A1-A3).\nAssuming that /tildewideΓ>>4γ, it makes sense to consider the\nbehavior at higher frequencies Ω ≫(ωf,c,γ) (regime I)\nand low frequencies (regime II) separately. In regime I\none finds:\nχ+−\nimp(Ω) =2µ2\nB(gf−gc)2\nπ/tildewideΓi/tildewideΓ\nΩ−ωf+i/tildewideΓ.(6)\nNeutron scattering data on diluted magnetic alloys show\na broad resonance in χ+−\nimp(Ω) of width TK5, in accor-\ndance with the above result. Note that this broad peak\nvanishes in the case of equal g-factors, as a consequence\nof the conservationof magnetization in that case (leaving\naside spin lattice relaxation for the moment). The result\nin regime II is:\nχ+−\nimp(Ω) =2µ2\nB\nπ/tildewideΓ{(gf−gc)2+gc(3gf−2gc)−ωc\nΩ−ωc+4iγ\n+g2\nc−ωc(ωf−ωc)\n(Ω−ωc+4iγ)2} (7)\nThe last term carries vanishing spectral weight. The sec-\nond term on the r.h.s. represents an impurity induced\nenhancement (3 gf>2gc) or reduction of the bulk con-\nduction electron spin resonance. This contribution van-\nishes ifgf=2\n3gc. The static susceptibility takes the form\nχ+−\nimp(0) = 2µ2\nBg2\nf/(π/tildewideΓ) .\nTo summarize, the impurity induced component of the\ntotaldynamicalspinsusceptibilityofaKondoionischar-\nacterized by a broad excitation peak of width /tildewideΓ≃TKat\nΩ =ωfand a narrow peak or dip of width 4 γat Ω =ωc,\nwhereγis the conduction electron relaxation rate. The\nrelativeweightsofthetwocomponentsdependsensitively\non the ratio of g-factors. This structure is not easily de-\ntected in an ESR experiment. The narrow resonance line\nhas the same position and width as the bulk ESR line.\nIts weight per atom is, however, enhanced by the large\nfactorTF/TK, which comes from the renormalized sus-\nceptibility scale prefactor ∝1//tildewideΓ. Therefore, the ESR\nresponse of a diluted magnetic alloy with a concentra-\ntion of Kondo ions c > T K/TFwill be dominated by\nthe impurity contribution determined in the above Eqs.\n(6,7).III. ANDERSON LATTICE MODEL IN THE\nKONDO SCREENED REGIME.\nThe Hamiltonian of the simplest Anderson lattice\nmodel, assuming momentum independent hybridization\nand an isotropic conduction band with flat density of\nstates is given by\nH=Hc+/summationdisplay\ni,σǫfσf+\niσfiσ+U/summationdisplay\ninfi↑nfi↓\n+V/summationdisplay\ni,k,σ(eik·Rif+\niσckσ+h.c.) (8)\nHereHcandǫfσ,V, Uhave been introduced in Sec. II\nand the Riare lattice site vectors. The single particle\nGreen’s functions are given by Dyson’s equation\n/parenleftbigg\niωn−ǫfσ−Σfσ(iωn,k) −V\n−V iω n−ǫkσ−Σcσ(iωn,k)/parenrightbigg\nG= 1,\nwhere G=/parenleftbigg\nGff\nkσGcf\nkσ\nGfc\nkσGcc\nkσ/parenrightbigg\n. (9)\nWe assume Fermi liquid theory to hold. Then the self\nenergy Σ fσ(ω,k) has a power series expansion in ωat\nthe Fermi surface, and its imaginary part is small ∝ω2,\nand may be neglected in lowest order. One may use\nω−ǫfσ−Σfσ(ω,kF) =z−1\nσ[ω−/tildewideǫfσ], with the quasipar-\nticle weight factor zσ= [1−(∂Σfσ(ω,kF)/∂ω)0]−1and\nthe renormalized energy /tildewideǫfσ=zσ[ǫdσ+Σfσ(0,kF)]. The\nconduction electron self-energy may be approximated by\nΣcσ(ω+i0,k) =−iγ, whereγis the spin-lattice relax-\nation rate defined earlier. Then for low energies one\nhas a quasiparticle description, with Gff\nkσ(ω) =zσ/tildewideGff\nkσ,\nGcf\nkσ=√zσ/tildewideGcf\nkσand the renormalized hybridization\namplitude /tildewideV2=zσV2. The full matrix of quasiparticle\nGreen’s functions is given by\n/parenleftbigg/tildewideGff\nkσ/tildewideGcf\nkσ/tildewideGfc\nkσGcc\nkσ/parenrightbigg\n=1\ndet/parenleftbigg\nω−ǫkσ+iγ/tildewideV\n/tildewideV ω −/tildewideǫfσ/parenrightbigg\n,(10)\nwhere det = ( ω−/tildewideǫfσ)(ω−ǫkσ+iγ)−/tildewideV2= (ω−ζ+\nkσ)(ω−\nζ−\nkσ). The complex energy eigenvalues are given by\nζ±\nkσ=1\n2(/tildewideǫfσ+ǫkσ−iγ)±/radicalbigg\n1\n4(/tildewideǫfσ−ǫkσ+iγ)2+/tildewideV2\n=ǫ±\nkσ−iγ±\nkσ, (11)\nwhereǫ±\nkσ= Reζ±\nkσandγ±\nkσ=−Imζ±\nkσ. We note γ±\nkσ>\n0. There are two energy bands separatedby an (indirect)\ngap (ǫmin,max\nkσdenote the minimum or maximum of the\nconduction band):\n∆gσ=ǫmin\nkσ−ǫmax\nkσ+/radicalBig\n(/tildewideǫfσ−ǫmax\nkσ)2+4/tildewideV2\n+/radicalBig\n(/tildewideǫfσ−ǫmin\nkσ)2+4/tildewideV2≫ωf,c.(12)\nWe assume that the renormalized f-level/tildewideǫfσis inside the\nconduction band, close to the Fermi level, and consider4\nthe case of almost half-filling, i.e. n/lessorsimilar2 electrons per\nlattice site. Then only the lower band is occupied in the\nground state. We assume an isotropic band structure for\nsimplicity. Then near the Fermi level at k=kF, the\nquasiparticle energy (we drop the spin dependence) has\nthe form\nǫ−\nk−ǫ−\nkF=1\n2(k−kF)vF\n1+(/tildewideǫf−ǫkF)/radicalBig\n(/tildewideǫf−ǫkF)2+4/tildewideV2\n\n≃(k−kF)v∗\nF, (13)\nwhere the renormalizedFermi velocity is defined by v∗\nF=\nvF/tildewideV2/(/tildewideǫf−ǫkF)2=vF(m/m∗)≃vFzV2/(ǫkF)2, and we\nused the fact that ǫkF≫/tildewideǫf. Note that ǫkFis the bare\nconduction band energy at k=kF, which is far above\nthe true Fermi energy. When z≪1, the Fermi velocity\nis renormalized to very low values and one has a “heavy\nfermion liquid” (effective mass m∗≫m). The effective\nFermi temperature of the heavy quasiparticles is given\nbyT∗\nF=1\n2kFv∗\nFσ≪TF.\nTo first order in γthe level widths are given by\nγ±\nkσ=1\n2γ/bracketleftbigg\n1∓/tildewideǫfσ−ǫkσ\nǫ+\nkσ−ǫ−\nkσ/bracketrightbigg\n. (14)\nWe note that the hybridization induced width /tildewideΓ of\nthef-electron energy level in the impurity problem is\nnow absorbed in the electronic band structure: the co-\nherent superposition of contributions from all lattice\nsites to Σ fσ(ω+i0,k) removes the large constant i/tildewideΓ.\nThe remaining imaginary part of the self-energy at fi-\nnite temperatures may be approximated by a constant\nΓFL=cT2/T∗\nF. We shall comment on the effect of quasi-\nparticle scattering on the ESR-linewidth at the end. Us-\ning partial fraction decomposition, we get the retarded\nGreen function\n/tildewideGff\nkσ(ω+i0) =aff,+\nkσ\nω−ζ+\nkσ+aff,−\nkσ\nω−ζ−\nkσ(15)\nand similar expressions for /tildewideGcf\nkσandGcc\nkσ, where, with\nukσ=ζ+\nkσ−ζ−\nkσ,\naff,±\nkσ=±(ζ±\nkσ−/tildewideǫkσ)/ukσ,\nacc,±\nkσ=±(ζ±\nkσ−ǫfσ)/ukσ,\nacf,±\nkσ=±/tildewideV/ukσ.\nFor sufficiently small spin-lattice relaxation, γ≪\n(/tildewideV,/tildewideǫfσ), we may neglect the imaginary parts in the\nweight factors aff,±\nkσ,...and replace ζ±\nkσbyǫ±\nkσ.\nThe quasiparticles interact via the Fermi liquid inter-\naction. For ESR, the relevant component of the Fermi\nliquidinteractionisthe isotropicspin-antisymmetricpart\ndescribed by the Landau parameter Fa\n0. An important\ncontribution to Fa\n0comes from the renormalized value\n/tildewideUof the bare interaction U, Fa\n0=−2N0/tildewideU, similar tothe single impurity case discussed in Sec. II. For the lat-\ntice case, exact numerical results on /tildewideUare not available.\nWe note however, that the onsite repulsion Uis likely\nto be screened down to a positive value of order N−1\n0,\nwhich would lead to a ferromagnetic Landau parameter\n0> Fa\n0/greaterorsimilar−1. Additional contributions to Fa\n0may be\ngenerated by nonlocal interactions like the RKKY inter-\naction, which may be ferro- or antiferromagnetic. We\nemphasize that this Fermi liquid interaction always leads\nto a ferromagnetic contribution to the fluctuation spec-\ntrum, which may be more or less important depending\nupon the other contributions.\nFollowing the way in which the interaction was in-\ncluded in the impurity model, we may express the fully\nscreened f-electron susceptibility in terms of the un-\nscreened one\nχ+−\nff(iΩm) =χ+−\nff,H(iΩm)/[1−/tildewideUχ+−\nff,H(iΩm)],(16)\nwhere\nχ+−\nff,H(iΩm) =−T/summationdisplay\nωn/summationdisplay\nk/tildewideGff\nk↓,H(iωn+iΩm)/tildewideGff\nk↑,H(iωn)\n(17)\nTheone-to-onecorrespondenceofquasiparticlesandbare\nparticles, on which Landau’s Fermi liquid theory rests,\nallows to calculate the spin susceptibility from the quasi-\nparticle Green’s functions defined above, without taking\nthe incoherent parts into account. Here the subscript H\nindicates that the Zeeman energy ωfis replaced every-\nwhere by\n/tildewideωf=ωf[1+/tildewideUχ+−\nff(0)] =ωf[1−/tildewideUχ+−\nff,H(0)]−1(18)\nUsing the representation of /tildewideGff\nk↓,Hin terms of eigen-\nstates, the summation in Eq. (17) on ωnandkmay be\ndone. In the case that only the lower band is occupied,\nthe low frequency response is given by, see the Appendix,\nEq. (A4):\nχ+−(Ω+i0) =χ+−(0)(−ωr+iγr)\nΩ−ωr+iγr,(19)\nwhereχ+−(0) is defined in the Appendix, Eq. (A5). The\nmean field shift largely cancels out of the resonance fre-\nquency\nωr=1\n2ωf\n1−(/tildewideǫf−ǫkF)/radicalBig\n(/tildewideǫf−ǫkF)2+4/tildewideV2\n≃ωf(1−m/m∗),\n(20)\nThe linewidth, however, is reduced by the exchange in-\nteraction, provided the interaction is ferromagnetic.\nγr=γ\n1+/tildewideǫf−ǫkF/radicalBig\n(/tildewideǫf−ǫkF)2+4/tildewideV2\n[1−/tildewideUχ+−\nff,H(0)]\n≃2γm\nm∗[1−/tildewideUχ+−\nff,H(0)]≪γ. (21)5\nIt is seen that the main narrowingmechanism is provided\nby the hybridization throughthe renormalizedamplitude\n/tildewideV, which gives the small factor m/m∗. In simple terms,\nthe quasiparticles at the Fermi surface have mainly f-\ncharacter, with only a small admixture (fraction m/m∗)\nof conduction electron component. Since only the con-\nduction electronsfeel the spin lattice relaxation, the total\nspin relaxation is a fraction m/m∗of the spin lattice re-\nlaxation. Vertex correctionsto the spin-lattice relaxation\nare likely to increase γrsomewhat as they do in the im-\npurity case, Appendix Eq. (A1), where 2 γbecomes 4 γ.\nIn order to discuss the temperature and magnetic field\ndependence of the linewidth it is necessary to incorpo-\nrate quasiparticle scattering effects14and inelastic con-\ntributions to the spin-lattice relaxation. In the case that\ntheg-factors are sufficiently different, the contribution\nto the linewidth from quasiparticle scattering will vary\nwith temperature as T2/T∗\nFand with magnetic field H\nasH2/T∗\nF. In the case of equal g-factors the latter con-\ntribution will be cancelled by vertex corrections. Addi-\ntional temperature dependence may arise from coupling\nto phonons.\nIV. CONCLUSION\nThis paper is motivated by the recent observations\nof electron spin resonance at low temperature in some\nheavy-fermion compounds. We have calculated the dy-\nnamical susceptibility, which describes the resonance, at\nlow temperature in the fully screened Kondo regime for\nboth a single Kondo impurity spin as well as for the\nKondo lattice, described here by the Anderson lattice\nmodel.\nWe have not addressed the behavior of the susceptibil-\nities at temperatures in the neighborhood of the Kondo\ntemperature, where linewidths are expected to be very\nlarge due to rapid spin fluctuations in that temperature\nrange. Rather, we deal with the very low temperature\nregime, where a Kondo impurity is fully screened and\nwhere the heavy-electron Fermi liquid has formed in the\nAnderson lattice.\nFor the realistic case in which the g-factors of felec-\ntrons and conduction electrons are different, we find for\nthe single impurity that structure persists at both the f\nelectron and conduction electron resonance frequencies.\nThe impurity resonance continues to have a large width,\nof orderTK, while for the conduction electron resonance\nthere is an impurity-induced contribution that increases\nor decreases the amplitude depending on the ratio of g-\nfactors.\nThe situation is quite different for the lattice case.\nHere, the hybridization of the fand conduction elec-\ntrons and Fermi-liquid interaction lead to modifications\nof the susceptibility that can lead to substantial line nar-\nrowingand hence the possibility ofexperimentalobserva-\ntion. We find a sharp ESR line near the underlying local\nfelectron resonance. The line is substantially narrowedby a factor of the mass ratio m/m∗and by the effect of\nthe Fermi liquid interaction F0\naprovided it is negative\n(ferromagnetic).\nWenotethattheESRhasbeenonlybeenseeninheavy\nfermion compounds for which there is independent evi-\ndence for ferromagnetic fluctuations7,8. We suggest that\nour analysis accounts for this observation.\nAcknowledgments\nWe thank the Aspen Center for Physics, where this\nwork was begun and completed. We thank Q. Si for\npointing us to this problem and we acknowledge helpful\ndiscussions with K. Baberschke, P. Coleman, I. Martin,\nD. Pines. We are grateful to J. Sichelschmidt for his\ngenerous sharing of information and data. This work has\nbeensupportedinpartbytheDFG-CenterforFunctional\nNanostructurres and the DFG-Forschergruppe ”Quan-\ntum Phase Transitions” (PW).\nAPPENDIX\n1. Anderson impurity model in the Kondo screened\nregime: Green’s function approach to χ+−(Ω)\nAs derived in the main text, Eq. (2), the Green’s func-\ntions of conduction electrons and local electrons, includ-\ning the Fermi-liquid interaction, are given by\nGfσ(iωn) = [iωn+σωf+i/tildewideΓsign(ωn)]−1,\nGcσ(k,iωn) = [iωn−εk+σωc/2+iγsign(ωn)]−1.\nThe dynamical transverse susceptibility χ+−(Ω),\nwhere Ω is the frequency of an a.c. electromagnetic field\npolarized transverse to the static magnetic field, is given\nby\nχ+−(Ω) =µ2\nB[g2\ncχ+−\ncc(Ω)+g2\nfχ+−\nff(Ω)+2gcgfχ+−\ncf(Ω)].\nThe partial susceptibilities are obtained by evaluating\nFeynman bubble diagrams dressed by vertex corrections\noftheladdertypereferringtotheFermiliquidinteraction\n(local electrons) and the spin-orbit interaction (impurity\ncorrelation lines for the conduction electrons).\nThe local susceptibility in the absence of vertex cor-\nrections is obtained as\nχ+−\nff,H(iΩm) =−T/summationdisplay\nωnGf↓(iωn+iΩm)Gf↑(iωn)\n=2\nπ/tildewideΓ−ωf+i/tildewideΓ\niΩm−2ωf+2i/tildewideΓ.\nThe vertex corrections are obtained from the Bethe-\nSalpeter equation\nΛ(iΩm) = 1+/tildewideUχ+−\nff,H(iΩm)Λ(iΩm) =Ω−2ωf+2i/tildewideΓ\nΩ−ωf+i/tildewideΓ,6\nwhere we used /tildewideU=π/tildewideΓ. Then\nχ+−\nff(iΩm) =χ+−\nff,H(iΩm)Λ(iΩm) =2\nπ/tildewideΓ−ωf+i/tildewideΓ\niΩm−ωf+i/tildewideΓ.\nThe conduction electron susceptibility consists of four\ncontributions\nχ+−\ncc(iΩm) =χbulk\ncc(iΩm)+3/summationdisplay\ni=1χ(i)\ncc(iΩm)The bulk contribution has the form χbulk\ncc(iΩm) =\nNχ0\ncc(Ω +i0)Φ(iωn,iΩm), where Nis the num-\nber of atoms in the system and χ0\ncc(Ω +i0) =\n−T/summationtext\nωn,kGc↓(k,iωn+iΩm)Gc↑(k,iωn) =N0(−ωc+\n2iγ)/(Ω−ωc+2iγ), where N0is the conduction electron\ndensity of states at the Fermi level. The vertex function\nΦ(iωn,iΩm) is found as solution to the equation:\nΦ(iωn,iΩm) = 1−h2/summationdisplay\nkG0\nc↓(k,iωn+iΩm)G0\nc↑(k,iωn)Φ(iωn,iΩm)\nas\nΦ(iωn,iΩm) =θ(−ωn)θ(ωn+Ωm)iΩm−ωc+2iγ\niΩm−ωc+4iγ+[1−θ(−ωn)θ(ωn+Ωm)]\nNote that the minus sign in front of h2is generated by\nthe Pauli matrices that appear in Hc, Eq. (3) of the main\ntext (in the case ofpotential scatteringthere would be no\nsign change):/summationtext\ni,α,βσi\nα↓σi\n↑βG0\ncαG0\ncβ=−G0\nc↓G0\nc↑. As a\nconsequence, the vertexcorrectionsdouble the linewidth:\n2γ→4γ. In the case of potential scattering the vertex\ncorrections cancel the self-energy induced linewidth, so\nthat potential scattering does not contribute to the spinrelaxation, as expected. Combining the above results we\nfind\nχbulk\ncc(Ω+i0) =NN0−ωc+4iγ\nΩ−ωc+4iγ.(A.1)\nThe remaining contributions are obtained from\nχ(1)\ncc(iΩm) =−V2T/summationdisplay\nωn/summationdisplay\nk/braceleftBig\n[G0\nc↓(k,iωn+iΩm)]2G0\nc↑(k,iωn)Gf↓(iωn+iΩm)\n+G0\nc↓(k,iωn+iΩm)[G0\nc↑(k,iωn)]2Gf↑(iωn)/bracerightBig\n,\nχ(2)\ncc(iΩm) =−T/summationdisplay\nωn/bracketleftBigg\nV2/summationdisplay\nkG0\nc↓(k,iωn+iΩm)G0\nc↑(k,iωn)Φ(iωn,iΩm)/bracketrightBigg2\nGf↓(iωn+iΩm)Gf↑(iωn),\nχ(3)\ncc(iΩm) =−/bracketleftBigg\nV2T/summationdisplay\nωn/summationdisplay\nkG0\nc↓(k,iωn+iΩm)G0\nc↑(k,iωn)Φ(iωn,iΩm)Gf↓(iωn+iΩm)Gf↑(iωn)/bracketrightBigg2\n×[−/tildewideUΛ(iΩm)].\nUsing\n/summationdisplay\nk[G0\nc↓(k,iωn+iΩm)]2G0\nc↑(k,iωn) =N02πi\n(iΩm−ωc+2iγ)2=−/summationdisplay\nkG0\nc↓(k,iωn+iΩm)[G0\nc↑(k,iωn)]2\nand the identity\nGf↓(iωn+iΩm)−Gf↑(iωn) =−(iΩm−2ωf+2i/tildewideΓ)Gf↓(iωn+iΩm)Gf↑(iωn)\nas well as\nΠ(iΩm) =T/summationdisplay\n−Ωm<ωn<0Gf↓(iωn+iΩm)Gf↑(iωn) =1\nπ/tildewideΓiΩm\niΩm−2ωf+2i/tildewideΓ7\nwe get\nχ(1)\ncc(iΩm) =2\nπ/tildewideΓΩ\nΩ−2ωf+2i/tildewideΓ\nχ(2)\ncc(iΩm) =2\nπ/tildewideΓ2/tildewideΓ2\n(iΩm−ωc+4iγ)2Ω\nΩ−2ωf+2i/tildewideΓ\nχ(3)\ncc(iΩm) =−2\nπ/tildewideΓΩ2\n(iΩm−ωc+4iγ)2/tildewideΓ2\n(Ω−2ωf+2i/tildewideΓ)(Ω−2ωf+2i/tildewideΓ)\nAdding the three contributions we find\n3/summationdisplay\ni=1χ(i)\ncc(iΩm) =2\nπ/tildewideΓΩ(Ω−ωf)\n(iΩm−ωc+4iγ)2i/tildewideΓ\nΩ−ωf+i/tildewideΓ(A.2)\nThe mixed susceptibility may be calculated from the bubble diagram beg inning with a conduction electron particle-\nhole line and ending with a local electron p-h line, dressed by vertex co rrections at both ends:\nχ+−\ncf(Ωm) =−T/summationdisplay\nωn,kG0\nc↓(k,iωn+iΩm)G0\nc↑(k,iωn)Φ(iωn,iΩm)V2Gf↓(iωn+iΩm)Gf↑(iωn)Λ(iΩm)\n=2\nπ/tildewideΓ−i/tildewideΓ\niΩm−ωc+4iγiΩm\niΩm−2ωf+2i/tildewideΓ(A.3)\nAfter analytical continuation to the real frequency axis\nand combining the contributions the total impurity sus-\nceptibility is obtained as given in the main text, Eqs.\n(6,7).\n2. Anderson lattice model in the Kondo screened\nregime: Green’s function approach to χ+−(Ω).\nAs derived in the main text Eq. (10), the matrix of\nquasiparticle Green’s functions is given by\n/parenleftbigg/tildewideGff\nkσ/tildewideGcf\nkσ/tildewideGfc\nkσGcc\nkσ/parenrightbigg\n=1\ndet/parenleftbigg\nω−ǫkσ+iγ/tildewideV\n/tildewideV ω −/tildewideǫfσ/parenrightbigg\n,\nwhere det = ( ω−/tildewideǫfσ)(ω−ǫkσ+iγ)−/tildewideV2= (ω−ζ+\nkσ)(ω−\nζ−\nkσ). The complex energy eigenvalues are given by\nζ±\nkσ=1\n2(/tildewideǫfσ+ǫkσ−iγ)±/radicalbigg\n1\n4(/tildewideǫfσ−ǫkσ+iγ)2+/tildewideV2\n=ǫ±\nkσ−iγ±\nkσ,\nwhere, expanding to leading order in γ, as well as in\nωf,ωc,\nǫ±\nkσ= Reζ±\nkσ≃1\n2(/tildewideǫfσ+ǫkσ)±/radicalbigg\n1\n4(/tildewideǫfσ−ǫkσ)2+/tildewideV2\n≃ǫ±\nk−1\n2ω±\nkσ\nand\nγ±\nkσ=−Imζ±\nkσ≃1\n2γ[1±/tildewideǫfσ−ǫkσ\nǫ+\nkσ−ǫ−\nkσ]≃γ±\nk−1\n2η±\nkσ,with\nǫ±\nk=1\n2(/tildewideǫf+ǫk)±1\n2/radicalBig\n(/tildewideǫf−ǫk)2+4/tildewideV2\nω±\nk=1\n2(/tildewideωf+ωc)±1\n2/tildewideǫf−ǫk/radicalBig\n(/tildewideǫf−ǫk)2+4/tildewideV2(/tildewideωf−ωc)\nγ±\nk=1\n2γ[1∓/tildewideǫf−ǫk/radicalBig\n(/tildewideǫf−ǫk)2+4/tildewideV2]\nη±\nk=∓1\n2γ/tildewideωf−ωc/radicalBig\n(/tildewideǫf−ǫk)2+4/tildewideV24/tildewideV2\n(/tildewideǫf−ǫk)2+4/tildewideV2.\nThe susceptibility is, as in the case of the impurity,\ngiven by the sum of three contributions: ff,ccand\n(cf,fc):\nχ+−(Ω) =µ2\nB[g2\ncχ+−\ncc(Ω)+g2\nfχ+−\nff(Ω)+2gcgfχ+−\ncf(Ω)].\nHere the ff-susceptibility is screened by the Fermi liquid\ninteraction\nχ+−\nff(iΩm) =χ+−\nff,H(iΩm)Λ(iΩm),\nwith\nΛ(iΩm) = 1/[1−/tildewideUχ+−\nff,H(iΩm)]\nwhere\nχ+−\nff,H(iΩm) =−T/summationtext\nωn,k/tildewideGff\nk↓,H(iωn+iΩm)/tildewideGff\nk↑,H(iωn)8\nSimilarly,\nχ+−\ncf(iΩm) =χ+−\ncf,H(iΩm)Λ(iΩm),\nwhere\nchi+−\ncf,H(iΩm) =−T/summationtext\nωn,k/tildewideGcf\nk↓(iωn+iΩm)/tildewideGcf\nk↑(iωn)\nand\nχ+−\ncc(iΩm) =χ+−\ncc,H(iΩm)+/tildewideU[χ+−\ncd,H(iΩm)]2Λ(iΩm),\nwhere\nχ+−\ncc,H(iΩm) =−T/summationtext\nωn,kGcc\nk↓(iωn+iΩm)Gcc\nk↑(iωn)\nUsing the representation of the Green’s functions in\nterms of the eigenstates ν=±, and the fact that low\nenergyexcitationsareonlypossibleclosetothe Fermi en-\nergy, which we assume to lie in the lower band ( ν=−),\nonly the ( −)-components contribute to χ+−\nij,H(iΩm) :\nχ+−\nff,H(Ω+i0) =−/summationtext\nkaff,−\nk↓aff,−\nk↑f(ζ−\nk↑)−f(ζ−\nk↓)\nΩ−ζ−\nk↓+ζ−\nk↑+i0,\nwhere in the arguments of the Fermi function f(ǫ),\nthe complex-valued energy ζ−\nk↓appears. Employingζ−\nk↓−ζ−\nk↑≃ω−\nk−2iγ−\nk, and/summationtext\nk[f(ζ−\nk↑)−f(ζ−\nk↓)]≃\n/summationtext\nk(∂f/∂ǫ−\nk)(−ω−\nkF+2iγ−\nkF) =N0(ω−\nkF−2iγ−\nkF), we get\nχ+−\nff,H(Ω+i0) =N0aff,−\nkF↓aff,−\nkF↑−ω−\nkF+2iγ−\nkF\nΩ−ω−\nkF+2iγ−\nkF\nand hence\nχ+−\nff,H(0) =N0aff,−\nkF↓aff,��\nkF↑.\nEquivalent expressions hold for the ffandcfcompo-\nnents. The vertex function follows as\nΛ(Ω+i0) =Ω−ω−\nkF+2iγ−\nkF\nΩ−(ω−\nkF−2iγ−\nkF)(1−/tildewideUχ+−\nff,H(0))\nand the renormalized ff-susceptibility takes the form\nχ+−\nff(Ω+i0) =χ+−\nff(0)−ωr+iγr\nΩ−ωr+iγr,where\nωr−iγr= (ω−\nkF−2iγ−\nkF)(1−/tildewideUχ+−\nff,H(0)) and\nχ+−\nff(0) =χ+−\nff,H(0)/[1−/tildewideUχ+−\nff,H(0)]\nas discussed in the main text, Eqs. (20,21). The total\nsusceptibility consists of two resonance terms:\nχ+−(Ω+i0) =χ+−\nr(0)−ωr+iγr\nΩ−ωr+iγr+g2\nc−ω−\nkF+2iγ−\nkF\nΩ−ω−\nkF+2iγ−\nkF{χ+−\ncc,H(0)+/tildewideU[χ+−\ncf,H(0)]2−ωr+iγr\nΩ−ωr+iγr},(A.4)\nwhereχ+−\nr(0) =g2\nfχ+−\nff(0)+2gcgfχ+−\ncf(0). In the case that γ−\nkF≫γr, the resonance part simplifies to\nχ+−(Ω+i0) =χ+−(0)−ωr+iγr\nΩ−ωr+iγr,whereχ+−(0) =χ+−\nr(0)+g2\nc/tildewideU[χ+−\ncf,H(0)]2(A.5)\n1A.C. Hewson, The Kondo Problem to Heavy Fermions .\n(Cambridge University Press, 1993)\n2J. Kondo, Prog.Theor. Phys. 32, 37 (1964).\n3K. G. Wilson, Rev. Mod. Phys. 47, 773 (1975).\n4For references to the early experimental and theoretical\npapers, see W. Brenig, W. G¨ otze and P. W¨ olfle, Phys.\nRev. B2, 4533 (1970); J. Sweer, D.C. Langreth and J.W.\nWilkins, Phys. Rev B 13, 192 (1976).\n5U. Walter, E. Holland-Moritz and Z. Fisk, Phys Rev B 43,\n320 (1991)\n6J. Sichelschmidt, V.A. Ivanshin, J. Ferstl, C. Geibel and\nF. Steglich, Phys. Rev. Lett. 91, 156401-1 (2003).\n7C. Krellner, T. Forster, H. Jeevan, C. Geibel and J.\nSichelschmidt, Phys. Rev. Lett. 100, 066401 (2008).8Ref. 6 and U. Schaufuß, V. Kataev, A.A. Zvyagin, B.\nB¨ uchner, J. Sichelschmidt, J. Wykhoff, C. Krellner, C.\nGeibel, and F. Steglich, arXiv0804.4105v1. Also, J.G.S.\nDuque, E.M. Bittar, C. Adriano, C. Giles, L.M. Holanda,\nR. Lora-Serrano, P.G. Pagliuso, C. Rettori, C.A. P´ erez,\nR.W. Hu, C. Petrovic, S. Maquilon, Z. Fisk, D.L. Huber,\nand S.B. Oseroff, preprint.\n9George Feher and A.F. Kip, Phys. Rev. 98, 337 (1955).\n10E. Abrahams, unpublished.\n11P. Nozi` eres, J. Low Temp. Phys. 17, 31 (1974).\n12A.C. Hewson, J. Bauer and W. Koller, Phys. Rev. B 73,\n045117 (2006).\n13S.E. Barnes and J. Zitkova-Wilcox, Phys. Rev. B 7, 2163\n(1973).9\n14A.W. Overhauser, Phys. Rev. 89, 689 (1953)." }, { "title": "1702.05389v2.Charge_Transport_and_Entropy_Production_Rate_in_Magnetically_Active_Molecular_Dimer.pdf", "content": "Charge Transport and Entropy Production Rate in Magnetically Active Molecular Dimer\nJ. D. Vasquez Jaramillo and J. Fransson\u0003\nDepartment of Physics and Astronomy, Box 516, 75120, Uppsala University, Uppsala, Sweden\n(Dated: October 23, 2018)\nWe consider charge and thermal transport properties of magnetically active paramagnetic molecular dimer.\nGeneric properties for both transport quantities are reduced currents in the ferro- and anti-ferromagnetic regimes\ncompared to the paramagnetic and e \u000ecient current blockade in the anti-ferromagnetic regime. In contrast, while\nthe charge current is about an order of magnitude larger in the ferromagnetic regime, compared to the anti-\nferromagnetic, the thermal current is e \u000eciently blockaded there as well. This disparate behavior of the thermal\ncurrent is attributed to current resonances in the ferromagnetic regime which counteract the thermal flow. The\ntemperature di \u000berence strongly reduces the exchange interaction and tends to destroy the magnetic control of the\ntransport properties. The weakened exchange interaction opens up a possibility to tune the system into thermal\nrectification, for both the charge and thermal currents.\nI. INTRODUCTION\nThermal transport properties in molecular junctions can be\nof electronic origin or mediated through lattice vibrations [1].\nThere has been an increasing interest in the study of ther-\nmal properties in molecular junctions [2–5], stimulated by\nexperimental observations [2]. Several realizations of tun-\nneling junctions comprising noble metal electrodes and poly-\nmers that absorb, emit, and transmit thermal have been re-\nported [4, 6]. The interest is, moreover, driven from the\nperspective of information science and technology with re-\nspect to entropy production rate [3, 7] and the meaning of the\nthermodynamics in low dimensional systems [3, 8, 9]. This\nwas motivated by the discovery of conducting polymers and\nsolitonic electronic transport mechanisms discovered by Shi-\nrakawa, see [10] and references therein. By this discovery\npolyacetylene became the test bench, bridging the gap be-\ntween organic and inorganic chemistry regarding electronic\ntransport [11–13]. Since then, charge transport has been ex-\ntensively studied theoretically and experimentally in molec-\nular junctions [15–25]. From the same standpoint, thermal\ntransport studies were conceived and came to be conclusive\nin the upcoming years [2]. Subsequently, several theoretical\nstudies demonstrated the possibility to conduct both thermal\nand charge through tunneling junctions [26–34], both in pres-\nence and absence of lattice vibrations, although there is no\ngeneric framework that successfully is capable of describing\nthe thermodynamic properties of nano junctions [8, 14, 35].\nTheoretical predictions suggest all electrical control for\nboth reading and writing spin states in molecular dimers\n\u0003Electronic address: Jonas.Fransson@physics.uu.se[51, 52]. This prediction is based on the (electronically me-\ndiated) indirect exchange interaction between the localized\nspins which controls the charge transport properties [51, 52].\nSimilar e \u000bects were reported in Ref. [6], where the spin\nground state of a single metal complex is electrically con-\ntrolled, imposing transition between high ( S=5=2) and low\n(S=1=2) spin configurations in a three terminal device.\nHere we build on the previous predictions made in Ref.\n[51, 53] for dimers of magnetic molecules in which the ef-\nfective spin-spin interactions are mediated by the properties\nof the delocalized electrons and extend to thermally induced\nmagnetic and transport properties. In particular we study\nthermal transport and its response to changes of the mag-\nnetic configurations. Our set-up pertains to, for instance,\nM-phthalocyanine (MPc), M-porphyrins, where M denotes a\ntransition metal atom [61–64], e.g., Cr, Mn, Fe, Co, Ni, Cu,\nand also to bis(phthalocyaninato)R (TPc 2) [65, 66], where R\ndenotes a rare earth element, e.g., Tb. Such molecules can be\ninvestigated in, for example, mechanically controlled break-\njunctions [66, 67], in carbon nanotube assemblies [65] and\nscanning tunneling microscope [62, 63].\nWe consider thermal and charge transport as the result of\nthe electrothermal control of the junction. Accordingly, we in-\nvestigate the charge and thermal conductance with respect to\nthe bias voltage and thermal gradient across the junction. With\nthis background we, furthermore, consider a non-equilibrium\nanalogue to the Seebeck coe \u000ecient defined as the ratio be-\ntween the di \u000berential conductances with respect to the ther-\nmal gradient and bias voltage, introduced in [68]. By the\nsame token, we consider the ratio of the energy di \u000berentials\nwith respect to the thermal gradient and the bias voltage. Our\npredictions and results are based on non-equilibrium Green\nfunctions defined on the Keldysh contour.\nThroughout this study we consider spin degenerate con-arXiv:1702.05389v2 [cond-mat.mes-hall] 23 Jan 20182\nditions in that we assume non-magnetic metals in the leads\nas well as the absence of externally applied magnetic fields.\nThe advantage with this set-up, compared to designs based\non ferromagnetic leads is that the dipolar and quadrupolar\nfields considered in Ref. [69] here becomes vanishingly\nsmall. Therefore the e \u000bective isotropic electron mediated\nspin-spin interactions dominates the properties and control of\nthe molecular dimer. In this sense, our system would serve as\na representation of paramagnetic spintronics, or paratronics .\nII. METHODS\nA. Magnetic molecular dimer in a junction\nThe specific set-up we address in this article comprises a\ndimer of paramagnetic molecules which are embedded in se-\nries in the junction between normal metallic leads, see Figure\n1 (a). Each paramagnetic molecule comprises a localized spin\nmoment which is embedded in a ligand structure, where the\nsp-orbitals define the spin-degenerate HOMO or LUMO or-\nbitals. We assume that the d-, or, f-orbitals that constitute the\nmolecular spins, hybridize only weakly with the sp-orbitals,\nallowing to consider the spin moment in the localized moment\npicture. As such, the localized moment interacts with the de-\nlocalized electrons only via local exchange. We also neglect\nspin-orbit coupling in the molecular orbitals as well as consid-\nering them in single electron form, which is typically justified\nfor the sp-electrons. The molecular orbitals couple via tun-\nneling both to one another and to the adjacent lead.\nWe model this set-up using the Hamiltonian\nH=HM+Hint+HL+HR+HT: (1)\nHere, the molecular HOMO or LUMO levels are defined by\nHM=X\n\u001b0BBBBBB@X\nm=L;R\"mdy\nm\u001bdm\u001b+Tc(dy\nL\u001bdR\u001b+H:c:)1CCCCCCA;(2)\nwhere dy\nm\u001b(dm\u001b) creates (annihilates) an electron in the\nleft ( L) or right ( R) molecule at the energy \"m=\"0and\nspin\u001b=\";#, whereasTcdefines the tunneling rate be-\ntween the molecules. Internally in molecule m, the local-\nized spin moment Sminteracts with the electron spin sm=P\n\u001b\u001b0dy\nm\u001b\u001b\u001b\u001b0dm\u001b0=2, where \u001bis the vector of Pauli matri-\nces, via exchange\nHint=X\nm=L;Rvmsm\u0001Sm; (3)\nwhere vmis the exchange integral, and we assume that vm=v.\nWe focus on the case with non-magnetic leads,\nHL=R=X\nk\u001b2L=R\"kcy\nk\u001bck\u001b; (4)\nwhere cy\nk\u001bcreates an electron in the left ( L;k=p) or right ( R;\nk=q) lead at the energy \"kand spin\u001b. Tunneling between\nFIG. 1: (Color online) (a) Molecular dimer of paramagnetic\nmolecules. An electron (at energy \"0) in each molecule interacts\nwith the localized spin moment ( Sm,m=L;R) via exchange ( vm)\nwith the electron in the adjacent molecule (tunneling rate Tc) and\nwith electron sin the left /right electrode (coupling \u0000). The left /right\nnonmagnetic electrode is characterized by its electrochemical poten-\ntial (\u0016L=R). E\u000bective molecular orbutals ( \"0\u0006Tc) emerge from in-\ntermolecular tunneling. (b) E \u000bective exchange interaction between\nthe localized spin moments as function of the voltage bias V. (c)\nOccupation of the states in the spin dimer. The green curve repre-\nsents the occupation of the lowest energy eigenstate of the spin dimer\nwhich changes character between spin singlet and spin triplet states\nas function of the voltage bias. Other colors analogously represent\nthe occupation of the consecutively higher energy eigenstates. In the\nregion indicated by the red arrow three states are degenerate and form\na spin triplet. Calculations are made at TL=4 K and TR=9 K for\n\"0\u0000\u0016=1,Tc=1 meV , vm=5 meV , and \u0000 =1, meV . In panels (b)\nand (c), the ferromagnetic and antiferromagnetic regimes of the spin\ndimer are indicated with red and black arrows, respectively.\nthe leads and molecules is described by\nHT=X\np\u001bTLcy\np\u001bdL\u001b+X\nq\u001bTRcy\nq\u001bdR\u001b+H:c: (5)\nand we define the voltage bias Vacross the junction by eV=\n\u0016L\u0000\u0016R, where\u0016\u001f,\u001f=L;R, denotes the electrochemical po-\ntential of the lead \u001f.\nIn this wayH0=HL+HR+HT+HMprovides a spin-\ndegenerate background electronic structure which mediates\nthe exchange interactions between the localized spin moments\ninHint. The spin-degeneracy implies that these interactions\nare purely isotropic [51, 53], such that we retain the Heisen-\nberg model only for the spins.\n1. Exchange interactions\nIt has been shown that the Hamiltonian in Eq. (1) gives rise\nto the e \u000bective spin model [53]\nHS=X\nmn\u0010\nJmnSm\u0001Sn+Dmn\u0001[Sm\u0002Sn]+Sm\u0001Imn\u0001Sn\u0011\n:(6)\nHere, the parameter Jmndenotes the isotropic Heisenberg\ninteraction, whereas DmnandImnrespectively denotes the\nDzyaloshinskii-Moriya and Ising interactions which both in-\ntroduce anisotropic interactions into the system. In the present3\nFIG. 2: (Color online) The exchange interaction J(units: meV) be-\ntween the molecular spins as function of voltage bias Vand temper-\nature di \u000berence \u0001T=TR\u0000TLfor di \u000berent gating conditions, such\nthat\"0\u0000\u0016=0;1;2;3 meV . Other parameters are as in Figure 1.\nsystem, where the there is no spin-dependence imposed, either\nby external or internal forces, the anisotropic interactions van-\nish,Dmn=0,Imn=0, and we retain the isotropic Heisenberg\ninteraction only.\nIn this paper, we treat the spin dimer as a closed system\nin the sense that we require a conserved number of particles\nfor which the occupations of the states are given by the Gibbs\ndistribution. This is a valid approximation for localized spins\nwhere the hybridization with the surrounding itinerant elec-\ntrons is small such that nature of the localized electrons can be\ndescribed in terms of a Kondo-like model rather than a fluc-\ntuating spin model in the sense of the Hubbard or Anderson\nmodels.\nThe spin-spin interactions are, nevertheless, influenced by\nthe tunneling current that flows through the molecular com-\nplex. In such set-ups, the e \u000bective magnetic interaction pa-\nrameter Jbetween the two spins can be calculated using the\nexpression, see Refs. [51, 53],\nJ=\u0000T2\nc\n8\u0019v2X\n\u001fZ\n\u0000\u001ff\u001f(!)(!\u0000\"0)(!\u0000\"0)2\u0000T2\nc\u0000(\u0000=8)2\nj(!\u0000\"0+i\u0000=8)2\u0000T2cj4d!;\n(7)\nwhere \u0000 =P\n\u001f=L;R\u0000\u001f, with \u0000\u001f=2\u0019P\nk\u001b2\u001fT2\n\u001f\u000e(!\u0000\"k), is the\ncoupling to the leads, and f\u001f(!) is the Fermi function at the\nchemical potential \u0016\u001f. Here, we have assumed that the molec-\nular level\"m=\"0,m=L;R, and that the local exchange cou-\npling vm=v, which is reasonable for equivalent molecules.\nWe remark here that for a more general molecular assembly,\nthat is, non-equivalent molecules and asymmetric couplings\nto the leads, the above expression for the exchange interac-\ntion should be replaced by the general formulas given in Ref.\n[53]. While this article is focused on dimers with equivalent\nmolecules coupled symmetrically to the leads, we briefly dis-\ncuss deviations from this case below.\nNotice that while the exchange interaction is defined\nin terms of the two-electron propagator ( \u0000i)hTs(t)s(t0)i\n[53], we employ the de-coupling approximation\n(\u0000i)sp\u001bG12(t;t0)\u001bG21(t0;t). Here, Gmn(t;t0) is a single-\nelectron Green function for the molecular orbitals projectedonto the sites mandn, whereas sp defines the trace over spin\n1=2 space. This approximation is justified when neglect-\ning local correlations; here the couplings to the localized\nspins. An obvious improvement would be to include these\ncorrelations and solve the two-electron propagator in a\nself-consistent scheme along with the Dyson equation for\nthe single electron Green functions. However, since such\nan approach unavoidably also would include self-consistent\ncalculations of the exchange and the spin configurations, and\nas we are mainly interested in qualitative e \u000bects, this is far\nbeyond the scope of the present article.\n2. Non-equilibrium variations\nThe salient features of the voltage bias dependence of Jand\nthe corresponding occupation of the spin states in the dimer\nare plotted in Figure 1 (b), (c) for the case with \"0\u0000\u0016=1 and\na temperature di \u000berence of 5 K (for other parameters, see the\nfigure caption). For the details about the interdependence be-\ntween the electronic structure in the molecular orbitals and the\nspin-spin exchange we refer to Refs. [51, 53]. The tempera-\ntures of the leads are included in the respective Fermi function\non which the exchange interaction within the spin dimer de-\npends, see Eq. (7). In addition to the dependencies on the\ntemperature of the leads, the exchange also depends on the\nenergy of the molecular levels. These dependencies are il-\nlustrated in Figure 2, which shows the exchange interaction\nenergy as function of the voltage bias and temperature di \u000ber-\nence for di \u000berent energies \"0\u0000\u0016. There is a clear distinction\nbetween the situation in panel (a) compared to panels (b) –\n(d), which originates in the level position relative to \u0016.\nFirst, whenever the localized level \"0\u0000\u0016=0, the molecu-\nlar bonding (at the energy: \"0\u0000T c) and anti-bonding (at the\nenergy:\"0+Tc) orbitals are centered symmetrically around\n\u0016which leads to that the influences from the left and right\nleads is symmetric with respect to voltage bias and indepen-\ndent of whether the source or drain electrode is warmer than\nthe other. The situation is schematically depicted in Figure 3\n(a), (b). The first thing to notice is a weakening of the ex-\nchange interaction caused by the increased thermal spread of\nthe electrons in the hot reservoir, compared to the cold reser-\nvoir. In equilibrium, then, in comparison to the cold reservoir\nthe hot reservoir contains a larger number of electrons above\nand below both molecular orbitals which contribute to an anti-\nferromagnetic interaction and simultaneously a smaller num-\nber of electrons that contribute to the ferromagnetic exchange.\nThis results in a weaker ferromagnetic interaction, which is\nverified in the simulations, see Figure 2 (a). The analo-\ngous argument holds for finite biases in the anti-ferromagnetic\nregime, however, utilized in the opposite way. Due to the\nthermal distribution of the electrons, there are electrons in\nthe hot reservoir contributing to both the ferromagnetic and\nanti-ferromagnetic exchange which results in an e \u000bectively\nweaker anti-ferromagnetic exchange, see Figure 2 (a). More-\nover, since the thermal distribution of the electrons is symmet-\nric around the pertinent chemical potential, the e \u000bect is equal\nregardless of the polarity of the voltage bias across the junc-4\ntion, which basically indicates voltage bias symmetric trans-\nport properties of the junction under these conditions.\nNext, whenever the molecule is gated such that the local-\nized level\"0is o\u000b-set from the chemical potential \u0016, a fi-\nnite temperature di \u000berence do generate an asymmetry with\nrespect to the voltage bias, see Figure 2 (b) – (d). One can\nconvince oneself that the exchange is symmetric with respect\nto the voltage bias at vanishing temperature di \u000berence, how-\never, in general the asymmetry is conspicuous. Under the\ngiven conditions, \"0\u0000\u0016>0 and variable TRwith TL=4 K\nfixed, the ferromagnetic regime at positive voltages remains\nnearly una \u000bected by changes in the temperature di \u000berence.\nThis can be schematically understood from the sketch in Fig-\nure 3 (c), showing \u0016Llying between the molecular orbitals.\nThen, the narrow thermal spread of the electrons in the left\nlead maintains a strong contribution to the ferromagnetic ex-\nchange while the contribution from the right lead is negligible\nmore or less independently of its temperature. By the same\ntoken one can understand that the ferromagnetic regime for\nnegative voltages is extremely sensitive to the temperature of\nthe right lead, see Figure 3 (d). Under these conditions, the\nexchange interaction is dominated by the contribution from\nthe right lead while the contribution from the left lead is neg-\nligible. Hence, at elevated temperatures there are competing\ncontributions to the exchange of both ferromagnetic and anti-\nferromagnetic nature which leads to a very weak ferromag-\nnetic exchange that weakens with increasing temperature dif-\nference.\n3. Deviations from perfectly equivalent and symmetric dimers\nIn realistic situations one can expect that the local parame-\nters vary from molecule to molecule, although they are meant\nto be considered as equivalent. Here, we briefly discuss some\nimplications on the indirect exchange interaction under di \u000ber-\nent local exchange integrals vm, finite level o \u000b-set\u0001 =\"L\u0000\"R,\nbetween the molecular levels \"m, and asymmetric couplings to\nthe leads.\nFor instance, for non-equivalent molecules such that \"1,\n\"2, the exchange interaction Jtends to become strongly asym-\nmetric with respect to the voltage bias, see for instance Ref.\n[51]. It may even lead to situations where Jis ferromag-\nnetic (negative) for one polarity of the voltage and anti-\nferromagnetic (positive) for the opposite. In those situations,\none would expect the resulting transport properties to be sig-\nnificantly di \u000berent for the two voltage polarities, something\nthat was also verified by a strong rectification property [51].\nExcept for the strong dependence of the exchange inter-\naction on the level o \u000b-set between the two molecules, the\ndimer structure is remarkably robust against small variations\nand asymmetries in the couplings \u0000\u001fto the leads as well as to\ninequalities in the local Kondo exchange couplings vm. Dif-\nferences of up to 20 % between the couplings of the two\nmolecules do not leads to any essential variations in the ex-\nchange interaction and are, therefore, not expected to be detri-\nmental to the transport properties either.\neV\n/uni03BCR/uni03BCLeV/uni03BCR\n/uni03BCL(a) (b) V>0 V<0 TL\nTLTRTR\neV\n/uni03BCR/uni03BCLeV/uni03BCR\n/uni03BCL(c) (d) V>0 V<0 TL\nTLTRTRFIG. 3: (Color online) The combined role of the voltage bias and\ntemperature di \u000berence on the tunneling conditions for (a), (c) pos-\nitive and (b), (d) negative bias, and (a), (b) symmetric and (c), (d)\nasymmetric molecular orbitals around \u0016(dashed). The left (right)\nlead is defined at the temperature TL(TR) and chemical potential \u0016L\n(\u0016R); in the figure TL\n\u001f(!)+f\u001f(\u0000!)G<\n\u001f(!)\u0011d!\n2\u0019;(11a)\nI\u001f\nN=(\u0000i)\u0000\u001f\n¯hspZ\u0010\nf\u001f(!)G>\n\u001f(!)+f\u001f(\u0000!)G<\n\u001f(!)\u0011d!\n2\u0019:(11b)\nHere, sp denotes the trace over spin 1 /2 space whereas G<=>\n\u001f,\n\u001f=L;R, denotes the 2\u00022-matrix Green function of the\nmolecule adjacent to the lead \u001f.\n1. Derivation of the molecular Green function\nWe make further analytical progress by constructing an ex-\nplicit expression for the Green function G\u001f. To this end we\ninclude the broadening e \u000bects from the couplings to the leads\nas well as the inter-molecular tunneling and intra-molecular\nexchange interactions with the localized spin moments. These\npresumptions lead to that we can write the retarded /advanced\nGreen functions G\u001fweighted on molecule \u001fas [51]\nGr=a\nL(R)(!)=1\n2˜TcX\ns=\u0006˜Tc\u001b0\u00002svhSz\nR(L)i\u001bz\n!\u0000Es\u0006i\u0000=8; (12)\nwhere the excitation energies E\u0006=\"0\u0006˜Tc=2 and ˜T2\nc=\nv2hSz\nL\u0000Sz\nRi2+4T2\nc.\nThe Green function Gfor the full dimer system is a 4 \u00024-\nmatrix G=fG\u001f\u001f0g\u001f;\u001f0=L;R, in which each entry is a 2 \u00022-matrix\nG\u001f\u001f0=fG\u001f\u001b\u001f\u001b0g\u001b;\u001b0=\";#. Here, the subscripts \u001f;\u001f0refer to the\nleft (right) molecule if \u001f\u001f0=LL(\u001f\u001f0=RR) and coupling be-\ntween the molecules for \u001f\u001f0=LRor\u001f\u001f0=RL. For brevity,\nwe write G\u001f\u001f=G\u001f. Each entry is defined by\nG\u001f\u001b\u001f0\u001b0(z)=hhd\u001f\u001bjdy\n\u001f0\u001b0ii(z)\n=Z\n(\u0000i)hTd\u001f\u001b(t)dy\n\u001f0\u001b0(t0)ie\u0000iz(t\u0000t0)dt0: (13)\nThe equation of motion for Gcan be summarized in the Dyson\nequation\nG=G0+G0\u0006G; (14)\nwhere G0is the bare Green function for the coupled\nmolecules, however, without couplings to the leads, whereas6\nFIG. 4: (Color online) (a) Molecule ( L,R) and spin projected (indi-\ncated by white arrows) densities of electron states of the left and right\nnon-gated (\u0016=0) molecules, calculated using GLL(RR)\u001bgiven in Eq.\n(16), respectively, as function of the voltage bias Vand energy !.\n(b) Molecule projected eq-fullGreenFunction of the localized spins\nhSz\n\u001fiand the total magnetic moment hSz\nL+Sz\nRi. Parameters are as in\nFigure 1.\u0006defines the self-energy generated by the couplings to\nthe leads. It can be noticed that since molecule 1 (2)\nonly couples to the left (right) lead, the retarded form of\nthis self-energy can be written as the diagonal matrix \u0006=\n(\u0000i)diagf\u0000L\n\";\u0000L\n#;\u0000R\n\";\u0000R\n#g=2. Considering spin-degenerate\nconditions, we can set \u0000\u001f\n\u001b= \u0000\u001f=2. As an e \u000bect of the Dyson\nequation for G, the corresponding lesser /greater forms are\ngiven by G<=>=Gr\u0006<=>Ga, where the lesser /greater forms\nof the self-energy is given by\n\u0006<=>(!)=(\u0006i)1\n4 \nfL(\u0006!)\u0000L\u001b00\n0 fR(\u0006!)\u0000R\u001b0!\n: (15)\nIn order to find the analytical forms of the local Green func-\ntion, we notice that since the spin degrees of freedom are\nuncoupled, we can write Gin block diagonal form. In this\nrepresentation, the blocks are distinguished by the spin index\nwhereas each block can be written on the form\nn\nGr=a\n\u001f\u001f0\u001b(!)o\n\u001f;\u001f0=L;R\n=\" \n!\u0000\"0\u0006i\u0000L=4\u0000Tc\n\u0000Tc!\u0000\"0\u0006i\u0000R=4!\n\u0000v\u001bz\n\u001b\u001b hSz\nLi0\n0hSz\nRi!#\u00001\n= !\u0000\"0\u0000vhSz\nRi\u0006i\u0000R=4Tc\nTc !\u0000\"0\u0000vhSz\nLi\u0006i\u0000L=4!\n(!\u0000Er=a\n+)(!\u0000Er=a\n\u0000);\n(16)\nwhere Er=a\ns=\"0+s˜Tc=2\u0007i(\u0000L+\u0000R)=8=\"0+s˜Tc=2\u0007i\u0000=8 and\nwhere we have put \u0000\u001f= \u0000=2 in the last equality. Accordingly,\nthe Green function weighted on the left molecule with spin \u001b\nis, therefore, given by the entry Gr=a\nLL\u001b(!), which can also be\nwritten on the form\nGr=a\nL\u001b(!)=1\n2˜TcX\ns=\u0006˜Tc\u00002svhSz\nRi\u001bz\n\u001b\u001b\n!\u0000Es\u0006i\u0000=8; (17)\nwith E\u0006=\"0\u0006˜Tc=2.\nWhen putting these results into the combination of the\nlesser and greater Green functions in the transport formulas,\nEq. (11), we retain only\nf\u001f(!)G>\n\u001f(!)\u0000f\u001f(\u0000!)G<\n\u001f(!)\n=(\u0000i)\u0000R\n2\u0010\nf\u001f(!)\u0000f\u001f0(!)\u0011\nGr\n\u001f\u001f0(!)Ga\n\u001f0\u001f(!);(18)\nout of which the only finite terms are the ones that couple the\ntwo molecules to one another and it is, therefore, necessary to\nstudy the forms of the site o \u000b-diagonal Green functions GLR\nandGRL. By a straight forward calculation one finds that theseGreen functions can be written as\nGr=a\n\u001f\u001f0(!)=Tc\n2X\ns=\u00061\u001b0\u0000s\u001bz\n!\u0000\"0\u0000svhSz\n\u001fi\u0006i\u0000=8Gr=a\n\u001f0(!)\n=Tc\n4˜TcX\nss0=\u00061[\u001b0\u0000s\u001bz][˜Tc\u001b0\u00002s0vhSz\n\u001fi\u001bz]\n[!\u0000\"0\u0000svhSz\n\u001fi\u0006i\u0000=8][!\u0000Es0\u0006i\u0000=8]:\n(19)\nIII. RESULTS AND DISCUSSION\nRegarding the charge transport across the junction, most of\nits properties can be understood in terms of the degree of de-\nlocalization of the electronic density in the molecular dimer.\nAs have been discussed in a previous publication [51], the\nmagnetic states and configurations lead to qualitatively dis-\ntinct regimes of the electronic properties of the dimer, some-\nthing that is conveyed over to the inherit transport properties\nof the molecular dimer. Accordingly, the conductance in the\nferromagnetic regime is expected to better, or, higher than in\nthe anti-ferromagnetic regime. This conjecture is based on the\nfact that in the anti-ferromagnetic regime the spin projected\nelectronic density in the bonding state is strongly localized to\noneorthe other molecule, see Figure 4 (a), and in the opposite7\n-6 -3\nvoltage bias (mV)3 6 002 IQ (MeV/s)\n-6 -3\nvoltage bias (mV)3 6 0-2×10-1(c) (d)-303IC (nA)\n-6 -3\nvoltage bias (mV)3 6 0 -6 -3\nvoltage bias (mV)3 6 0(a) (b)\nTR-TL=0 K\nTR-TL=10 K\nFIG. 5: (Color online) (a), (b), The charge current IC(units: nA)\nand (c), (d), entropy production rate IQ(units: MeV /s) as function\nof voltage bias VforTR\u0000TL=0 (black – faint) and TR\u0000TL=10\nK (red – bold) for (a), (c), \"0\u0000\u0016=0 meV and (b), (d), \"0\u0000\u0016=1\nmeV . Other parameters are as in Figure 1. In panels (a) and (b),\nthe ferromagnetic and anti-ferromagnetic regimes are indicated with\ngreen and blue arrows, respectively.\nfashion for the anti-bonding orbital. Therefore, an electron re-\nsiding in one of the molecules has a nearly vanishing probabil-\nity to tunnel over the other molecule which leads to a strongly\nsuppressed conductance. In the ferromagnetic regime, how-\never, the electronic density is more delocalized throughout the\nmolecular dimer which allows for resonant tunneling between\nthe molecules and, in turn, leads to an increased conductance.\nThese features and disparities of the two magnetic regimes\ncan be observed in the current, see Figure 5 (a), (b), showing\nthe charge current Icas function of the voltage bias at zero\ntemperature di \u000berence (black – faint) and TR\u0000TL=10 K (red\n– bold). Especially for vanishing temperature di \u000berence, the\nanti-ferromagnetic and ferromagnetic regimes (indicated by\nblue and green arrows, respectively) are strikingly separated\nby sharp current resonances which originates from a vanish-\ning exchange between the spin. The absence of the exchange\ninteraction leads to uncorrelated spins and a completely delo-\ncalized charge density in the molecular dimer, which allows\nfor a large current flow in the same way as in the high voltage\nregime where J!0.\nAt zero temperature di \u000berence across the junction, the\ncurrent-voltage characteristics is necessarily symmetric when-\never then electronic structure of the molecular dimer are sym-\nmetrically distributed, as is the case we consider here. In-\ntroduction of a finite temperature di \u000berence changes this sce-\nnario, however, a necessary condition for breaking the sym-\nmetric current-voltage characteristics is that the bonding and\nanti-bonding orbitals do not surround \u0016symmetrically. This\ncan be seen in the red – bold traces in Figure 5. In panel (a),\nthe molecular level \"0=\u0016which accordingly leads to a sym-\nmetric current as function of the voltage bias. In panel (b), in\ncontrast, the molecular level satisfies \"0\u0000\u0016=1 meV , which\nFIG. 6: (Color online) (a), (c), The charge current IC(units: nA)\nand (b), (d), entropy conductance IQ(units: MeV /s) as function of\nvoltage bias Vand temperature di \u000berence TR\u0000TLfor (a), (b), \"0\u0000\n\u0016=0 meV and (c), (d), \"0\u0000\u0016=1 meV . Other parameters are as in\nFigure 1\nthen generates a striking asymmetry in the current for finite\ntemperature di \u000berences. This result can be traced back to\nthe dramatically changed, weakened, exchange interaction be-\ntween the spins which causes an increased degree of delocal-\nization of the electronic density in the molecular dimer. In ef-\nfect, therefore, both the ferromagnetic and anti-ferromagnetic\nregimes are essentially destroyed for negative voltages, see\nFigure 5 (b) (red – bold). Hence, under the temperature di \u000ber-\nence between the leads the molecular dimer functions partially\nas a rectifying device where the magnetically active regimes\ncan be employed in the forward direction while the system\nbehaves like a normal conductor in the backward. In Fig-\nure 6 (a), (c), we show contour plots of the charge current\nas function of the voltage bias and temperature di \u000berence for\n(a)\"0\u0000\u0016=0 meV and (c) \"0\u0000\u0016=1 meV . The properties dis-\ncussed in detail above can be seen to be verified in the larger\npicture, varying continuously with the variations of the exter-\nnal conditions. Especially in the latter case ( \"0\u0000\u0016=1 meV),\nthe magnetically active regimes for negative voltage biases are\nseen to be destroyed already for small temperature di \u000berences,\nclosely following the behavior of the exchange interactions,\nc.f., Figure 2.\nNext, in the discussion of the entropy production rate, we\nagain notice that most of the expected features can be ex-\nplained in terms of the properties of the electronic structure\nin the molecular dimer analogous as with the charge current.\nHowever, for vanishing temperature di \u000berence between the\nleads a clear distinction compared to the charge current is that\nthe finiteness of the entropy production rate strongly depends\non the energy of the localized level in the molecular assem-\nbly. This can be traced back to the product of the energy and\nthe distribution functions ( G<=>\n\u001f) of the molecular dimer in,\nfor instance, Eq. (11a). One can make a simple compari-\nson between the qualitative properties of the two currents by\nassuming a Lorentzian model, 1 =[(!+\u0016\u0000\"0)2+(\u0000=2)2], for\nthe device embedded between the leads at low temperatures.\nThen, under the voltage bias \u0016L=R=\u0016\u0006eV=2, the two currents8\nbehave as\nIc\u00181\n\u0000=2\u0012\narctaneV=2\u0000\"0\n\u0000=2+arctaneV=2+\"0\n\u0000=2\u0013\n; (20a)\nIQ\u00181\n2ln(eV=2\u0000\"0)2+(\u0000=2)2\n(eV=2+\"0)2+(\u0000=2)2\n+\"0\u0000\u0016\n\u0000=2\u0012\narctaneV=2\u0000\"0\n\u0000=2+arctaneV=2+\"0\n\u0000=2\u0013\n:(20b)\nHence, while both the charge current and entropy production\nrate have normal on-sets associated with the energy \"0, de-\nscribed by the arctan-component, the entropy production rate\nis also logarithmically resonant at \"0. In addition, for the en-\ntropy production rate the on-set at \"0is weighted by the posi-\ntion of\"0relative to\u0016and, therefore, this contribution to the\nentropy production rate is expected to the small whenever \"0\nlies in the vicinity of \u0016. The logarithmic contribution tends to\nbe small for large voltage biases, since it leads to an increas-\ningly symmetric argument, and is accordingly only significant\nfor voltages such that either eV=2\u0000\"0oreV=2+\"0is small.\nExtrapolating this discussion to our present case with the\nmolecular dimer, we can verify these expected features under\nthe conditions of vanishing temperature di \u000berence. This can\nbe be seen in Figure 5 (c), (d) (black – faint). At \"0\u0000\u0016=0,\npanel (a), the entropy production rate is identically zero while\na finite o \u000b-set from\u0016, panel (b), yields a finite entropy pro-\nduction rate. In the latter case, the entropy production rate is\nsmall in the magnetically active regime and grows large only\nat large voltages where the exchange interaction vanishes. The\nentropy production rate tends to be e \u000eciently transmitted only\nwhen the molecular dimer is in its fully conjugated state, that\nis, when the exchange interaction is small. This is contrast to\nthe charge current where the less localized electronic density,\nin the molecular dimer, in the ferromagnetic regime yields a\nsignificant di \u000berence compared to the anti-ferromagnetic.\nThe reason for this qualitative di \u000berence between the\ncharge and entropy production rate, see Eq. (20), can be\nfound in the properties of the two currents in the ferromag-\nnetic regimes. In particular, the logarithmically resonant con-\ntribution to the entropy production rate, which is absent in the\ncharge current, peaks near the energies of the molecular bond-\ning and anti-bonding orbitals, something which occurs within\nthe ferromagnetic but not in the anti-ferromagnetic regime.\nThereto, these resonant peaks are oppositely directed com-\npared to the contributions varying like arctan x. As for the\nentropy production rate under finite temperature di \u000berence be-\ntween the leads, see Figure 5 (c), (d) (red – bold) and Figure\n6 (b), (d), this behavior is verified. This feature can be traced\nback to an increased degree of delocalization in the ferromag-\nnetic regime due to the increased thermal broadening from the\nhotter reservoir.\nIn the case where the localized level is o \u000b-set from\u0016, there\nis an interesting observation to be made in the temperature and\nvoltage dependence of both the charge and entropy production\nrate. At finite temperature di \u000berence and voltage bias, there\nis a strong asymmetry with respect to zero voltage which is\nclear signature that the magnetically active dimer potentially\ncan be used for current rectification. Previously it was shown\nT (K)048current (nA)\n-4 4 0\nT (K)-4 4 001020entropy prod. (MeV/s)\nT (K)-4 4 0012exchange (meV)(a) (b) (c)V=3 mV\nV=2.5 mV\nV=2 mV\nV=1.5 mV\nV=1 mV\nV=0.5 mVFIG. 7: (Color online) Thermal rectification. (a) Charge current, (b)\nentropy production rate, and (c) exchange interaction, as function of\nthe temperature di \u000berence for di \u000berent voltage biases. Here, \"0\u0000\u0016=\n2 andTc=3 meV , while other parameters are as in Figure 1.\nin Ref. [51] that the charge current can be rectified by intro-\nducing a level o \u000b-set between the localized levels in the two\nmolecules, something which possibly can be obtained by us-\ning di \u000berent types of molecules. The asymmetry induced by\nthe temperature di \u000berence provides a di \u000berent mechanism to\nrectification which is also viable for the entropy production\nrate. Indeed, the plots in Figure 6 (c), (d), illustrate that both\nthe charge and the entropy production rate is rectified at fi-\nnite temperature di \u000berences upon changing the polarity of the\nvoltage bias. For this to be successful, the dimer has to be\nset-up with a finite level o \u000b-set from\u0016and a finite voltage\nbias. Moreover, the parameters of the molecular dimer have\nto be tuned such that the o \u000b-set between the bonding and anti-\nbonding orbitals is larger than the thermal energy fed into the\ndimer from the hotter electrode. This enables a crossover from\nthe anti-ferromagnetic regime to either the ferromagnetic or\nparamagnetic regime, under changes in the temperature dif-\nference, which leads to strong variations in the currents. In\nprinciple, the stronger anti-ferromagnetic exchange one can\nobtain for one polarity of the temperature di \u000berence and the\nweaker the exchange can be in the opposite, regardless of sign,\nthe larger the ratio between the currents for the two polarities.\nIn Figure 7 we plot the (a) charge and (b) entropy produc-\ntion rate as function of the temperature di \u000berence for di \u000berent\nvoltage biases. We apply half the di \u000berence to the tempera-\nture in each lead such that TL=T\u0000\u0001=2 and TR=T+ \u0001=2,\nwhere the base temperature T=4 K. The calculations confirm\nthe argument that variations of the exchange interaction from\nstrongly anti-ferromagnetic to weakly ferromagnetic [ V\u00192\nmV , see Figure 2 (c)] indeed yields the larger ratio between\nthe large and small current regimes.\nFinally, we briefly discuss the di \u000berential conductances,\nboth charge conductance and di \u000berential entropy production\nrate and with respect to both voltage bias and temperature dif-\nference. Hence, we calculate @Ix=@Vand@Ix=@\u0001T, where\nx=c;Q, see Figs. 8 (a) – (d) and 9 (a) – (d). In Figs. 8\nand 9 we plot 1 =(1+expf\u001bFg), whereFis one of@Ix=@V,\n@Ix=@\u0001T, and Sx, in order to lower the values of the high am-\nplitudes. First we notice that both currents have ranges with\nfairly rapid variations with the voltage bias and with the tem-\nperature di \u000berence. While these properties can be traced back\nto the corresponding variations of the exchange interaction,\nwe can predict a few consequences of these features on the ra-9\nFIG. 8: (Color online) (a) – (d) Di \u000berential charge conductance (a),\n(c), and entropy production rate (b), (d), with respect to the voltage\nbias (a), (b), and temperature di \u000berence (c), (d). (e) Scand (f) SQ.\nHere,\"0\u0000\u0016=0, while other parameters are as in Figure 1. All con-\ntours show (1 +expf\u001bFg)\u00001, whereFis one of@Ix=@V,@Ix=@\u0001T,\nandSx, whereas\u001bis a scaling parameter.\ntios of the di \u000berential conductance with respect to the temper-\nature di \u000berence and the di \u000berential conductance with respect\nto the voltage bias, that is,\nSc=\u0000@Ic=d\u0001T\n@Ic=dV; (21)\nwhere the notation Scis used since the ratio coincides with the\nSeebeck coe \u000ecient in the limit V!0,\u0001T!0 [68]. In this\nsense the ratio Scis a non-equilibrium analogue of the See-\nbeck coe \u000ecient, however, we shall refrain from that nomen-\nclature for sake of not causing confusion with the concepts.\nNontheless, as rapid variations in the charge current yield\na large corresponding conductance and, oppositely, for slow\nvariations lead to a small conductance, we would expect that\nSctypically is large in the regions with weak variations of the\ncharge current. One is therefore led to think that Scmight\nbe large in the anti-ferromagnetic regime since the current is\nboth small and slowly varying with the voltage bias in those\nregimes, see Figs. 8 (a) and 9 (a). In addition, within the anti-\nferromagnetic regimes, @Ic=@\u0001Tvaries rapidly, including pos-\nsible sign changes, near zero temperature di \u000berence between\nthe leads. This conjecture is fairly well corroborated in the\nplots of the Scshown in Figs. 8 (e) and 9 (e). In particular, it\ncan be noticed in Figure 8 (e) that Scacquires large values in\nthe anti-ferromagnetic regime (voltages roughly in the ranges\n(\u00004;\u00002) and (2;4) mV) for temperature di \u000berences between\n0 K and 5 K. Except for these small regions of large Sc, it\ntends to be small in the remainder of the magnetically active\nregimes although not vanishing. In the case with a finite level\no\u000b-set, Figure 9 (e), however, Sctends to be large in the re-\ngions where the anti-ferromagnetic coupling is destroyed by\nFIG. 9: (Color online) (a) – (d) Di \u000berential charge conductance (a),\n(c), and entropy production rate (b), (d), with respect to the voltage\nbias (a), (b), and temperature di \u000berence (c), (d). (e), (f), Seebeck\ncoe\u000ecients associated with the (e) charge, Sc, and (f) heat, SQ, cur-\nrents. Here, \"0\u0000\u0016=1, while other parameters are as in Figure 1. All\ncontours show (1 +expf\u001bFg)\u00001, whereFis one of@Ix=@V,@Ix=@\u0001T,\nandSx, whereas\u001bis a scaling parameter.\nthe increased temperature di \u000berence. There are clearly visi-\nble domains at negative voltages which can be correlated with\nthe cross over between the anti-ferromagnetic and paramag-\nnetic regimes. At positive voltages where the dimer remains\nmagnetically active in a larger temperature range, Scis again\nsmall. We conclude that Scin both the ferromagnetic and anti-\nferromagnetic regimes is small and typically becomes finite at\nthe crossovers to the paramagnetic regime. Hence, the spin-\nspin interaction provides a way to not only control the charge\ncurrent but also Scin the system.\nIn analogy with the definition of Scassociated with the\ncharge current, we can define a coe \u000ecient SQfor the entropy\nproduction rate through\nSQ=\u0000@IQ=d\u0001T\n@IQ=dV: (22)\nAlthough we cannot give this coe \u000ecient an as simple interpre-\ntation as with the thermopower, we nevertheless find it useful\nin the analysis of the influence of the magnetic configurations\non the thermal properties. It can be seen in Figs. 8 (b) and\n9 (b) that the di \u000berential entropy conductance, with respect\nto the voltage bias, has a non-trivial dependence on the volt-\nage bias and temperature bias. Moreover, the dependence on\nthe temperature di \u000berence is intriguing. We can, however,\nnotice regarding the generic features of SQthat its more or\nless vanishing in the magnetically active regimes except in\nthe crossover between the di \u000berent regimes, where the en-\ntropy conductance varies slowly with the voltage bias but not\nnecessarily with the temperature di \u000berence. The qualitative\ndi\u000berence of the entropy conductance compared to the charge\ncurrent in the ferromagnetic regime, leads to that SQprovides10\na complimentary piece of information about the system under\ninvestigation in addition to Sc.\nIV . CONCLUSIONS\nIn summary we have theoretically studied the transport\nproperties of a dimer of paramagnetic molecules with local-\nized spins embedded in a junction between metallic leads. In\nparticular, we have addressed the charge and entropy conduc-\ntance flowing through the junction. It is demonstrated that\nthe indirect exchange interaction between the localized spins,\nwhich previously has been shown to depend on the voltage\nbias and temperature di \u000berence across the junction, acquires\na strongly asymmetric voltage bias dependence under finite\ntemperature di \u000berence between the leads. This property was\nsubsequently is predicted to have implications on, for exam-\nple, thermal rectification for both the charge and entropy con-\nductance. Simultaneously, our calculations suggest that the\ntemperature of the source electrode has a stronger influence\non the properties of the indirect exchange than the drain. It\nwas found, for instance, that while a voltage drop from the\nhotter to the colder reservoir tends to e \u000bectively destroy the\ntunable properties of the indirect exchange, these properties\nare stable under the opposite voltage polarity.\nThe transport properties of the dimer are intimately related\nto the indirect exchange, where the charge current is nearly\nblockaded for anti-ferromagnetic exchange, whereas it is fi-\nnite for ferromagnetic exchange, and maximal whenever the\nexchange is negligible. These three regimes can be explained\nby di \u000berent characteristics of the electronic density. In the\nanti-ferromagnetic regimes, the spin projections of the den-\nsity is strongly localized to one molecule such that transport\nbetween the molecules is suppressed. When in the ferro-\nmagnetic regime, the electron density is partially delocalized\nwhich leads to an enhanced conductance, whereas the cur-\nrent can flow freely in the paramagnetic regime due to a com-\npletely delocalized density. The entropy conductance follows\nthe same characteristics, as the charge current, in the anti-\nferromagnetic and paramagnetic regimes. In the ferromag-\nnetic regimes, however, the entropy conductance is strongly\nsuppressed by a contribution which is large, and nearly can-\ncels the regular entropy conductance contribution, near volt-\nage biases corresponding to the energy of the resonant states in\nthe molecular structure. By necessity, this resonant behavior\noccurs in the ferromagnetic regimes which leads to a strongly\nsuppressed entropy conductance there as well.\nWe, further, demonstrated that the non-equilibrium ther-\nmopower, in general is finite at the cross-over between\nregimes of di \u000berent indirect exchange associated with small\ndi\u000berential conductance. Typically, this behavior can be sum-\nmarized in that the thermopower is low within both the ferro-\nand anti-ferromagnetic regimes. Analogously, we introduced\na thermal coe \u000ecient as the ratio between the di \u000berential en-\ntropy conductance with respect to the temperature di \u000berence\nand voltage bias. While some of its properties closely fol-\nlow those ratios, additional features can be extracted, espe-\ncially at the cross over between the ferro- and antiferromag-netic regimes. In this sense, this ratio provides a complemen-\ntary sensitivity to the transport measurement which may show\nuseful in the analysis of the internal properties of the system.\nWe remark that while our results presented in this arti-\ncle are quite qualitative, they are realistic from the follow-\ning point of view. The results for the exchange interaction\nJpresented in the previous section, are obtained using the\nbare single electron Green functions for the molecular lev-\nels. This means that the back-action e \u000bect from the local\nspin moment is not included. In the calculations of the trans-\nport properties, on the other hand, the presence of the local\nspin moments are included, something which is crucial in\norder to investigate possible signatures in the transport data\nthat originates from the spin configurations. As for the trans-\nport calculations we could have chosen to simply demonstrate\nhow the charge and thermal transport depend on the nature\nof the exchange interaction, whether it is ferromagnetic (neg-\native), anti-ferromagnetic (positive), or paramagnetic (zero).\nGiven the approximation in which we calculate the local elec-\ntronic Green function of the molecular dimer, we would obtain\nthe transport characteristics that are presented in the Results.\nHowever, instead of making assumptions about the values of\nthe exchange interaction, we use the values as calculated by\nthe formulas provided in Exchange. In this way we incorpo-\nrate the voltage bias and thermal gradient dependence of the\nexchange also on the transport properties. While this approach\ncertainly has its limitations, we notice that a more thorough\nstudy of the correspondence between the regimes with dif-\nferent spin couplings and the associated transport properties\nrequires self-consistent calculations. Such calculations are,\nhowever, beyond the scope of the present investigation.\nConsidering the limitations of the method, we yet believe\nthat our findings are realistic and relevant to existing molecu-\nlar structures. The values of the local exchange interaction\nbetween localized spin and delocalized electrons can vary\nbetween 0.5 – 20 meV [72, 73], which leaves a large win-\ndow of tuning freedom with respect to couplings to the leads\nand HOMO /LUMO level o \u000b-set of the molecules. Moreover,\nsince our predictions are stable with respect to di \u000berences in\nthe local exchanges as well as the couplings to the leads, this\nalso allows for flexibility in the design of potential experi-\nments.\nThe predictions discussed in this paper opens the possibil-\nity to design nanoscale structures, in particular using mag-\nnetic molecules, that have a strong sensitivity on the local\nspin states of the system, which can be measured through the\ncharge and thermal transport characteristics. In ways, this sug-\ngests an alternative utilization of the spin degrees of freedom,\ncompared to the conventionally implemented spintronics, in\nwhich external magnetic fields are absent. The absence of\nsuch fields, in turn, leads to that the spin-spin interactions are\nisotropic which implies a truly magnetically isotropic (param-\nagnetic) device. Experimental confirmation of our predictions\nare feasible using state-of-the-art technology.11\nV . ACKNOWLEDGEMENT\nWe thank S. Borlenghi Garoia, A. Bouhon, K. Björnson, H.\nHammar, D. Kuzmanovksi and J. Schmidt for fruitful discus-sions. Financial support from Colciencias (Colombian Ad-\nministrative Department of Science, Technology and Innova-\ntion) and Vetenskapsrådet is acknowledged.\n[1] Huang, K. Statistical Mechanics ,1987 , 2nd ed. (Wiley).\n[2] Nitzan, A. Chemistry. Molecules Take the Heat. Science 2007 ,\n317, 759–760.\n[3] Pekola, J. P. Towards Quantum Thermodynamics in Electronic\nCircuits. Nature Phys. 2015 ,11, 118–123.\n[4] Reddy, P.; Jang, S.-Y .; Segalman, R. A.; Majumdar, A. Ther-\nmoelectricity in molecular junctions. Science 2007 ,315, 1568–\n1571.\n[5] Garrigues, A. R.; Wang, L.; del Barco, E.; Nijhuis, C. A.\nElectrostatic Control over Temperature-Dependent Tunnelling\nAcross a Single-Molecule Junction. Nat. Comm. 2016 ,7,\n11595.\n[6] Osorio, E. A.; Moth-Poulsen, K.; Van Der Zant, H. S. J.;\nPaaske, J.; Hedegård, P.; Flensberg, K.; Bendix, J.; Björnholm,\nT. Electrical Manipulation of Spin States in a Single Electro-\nstatically Gated Transition-Metal Complex. Nano Lett. 2010 ,\n10, 105–110.\n[7] Parrondo, J. M. R.; Horowitz, J. M.; Sagawa, T. Thermodynam-\nics of Information. Nat. Phys. 2015 ,11, 131–139.\n[8] Esposito, M.; Ochoa, M. A.; Galperin, M. Quantum Thermody-\nnamics: A Nonequilibrium Green’s Function Approach. Phys.\nRev. Lett. 2015 ,114, 080602.\n[9] Shastry, A.; Sta \u000bord, C. A. Temperature and V oltage Measure-\nment in Quantum Systems Far From Equilibrium. Phys. Rev. B:\nCondens. Matter 2016 ,94, 155433.\n[10] Dauxois, T.; Peyrard, M. Physics of Solitons 2006 (Cambridge\nUniversity Press, Cambridge).\n[11] Chiang, C.K.; Fincher, C.R.; Park, Y .W.; Heeger, A.J.; Shi-\nrakawa, H.; Louis, E. J.; Gau, S. C.; MacDiarmid, A. G. Electri-\ncal Conductivity in Doped Polyacetylene. Phys. Rev. Lett. 1977 ,\n39, 1098–1101.\n[12] Fincher, C. J.; Peebles, D.; Heeger, A.; Druy, M.; Matsumura,\nY .; MacDiarmid, A.; Shirakawa, H.; Ikeda, S. Anisotropic Op-\ntical Properties of Pure and Doped Polyacetylene. Sol. State\nComm. 1978 ,27, 489–494.\n[13] Su, W. P.; Schrie \u000ber, J. R.; Hegger, A. J. Soliton Excitations in\nPolyacetylene. Phys. Rev. B 1980 ,22, 2099–2111.\n[14] Ochoa, M. A.; Bruch, A.; Nitzan, A. Energy distribution and lo-\ncal fluctuations in strongly coupled open quantum systems: The\nextended resonant level model, Phys. Rev. B: Condens. Matter\nPhys. 2016 94, 035420.\n[15] Rammer, J.; Smith, H. Quantum Field-Theoretical Methods in\nTransport Theory of Metals. Rev. Mod. Phys. 1986 ,58, 323–\n359.\n[16] Heeger, A. J.; Kivelson, S.; Schrie \u000ber, J. R.; Su, W. P. Solitons\nin Conducting Polymers. Rev. Mod. Phys. 1988 ,60, 781–850.\n[17] Roth, S.; Bleier, H.; Pukacki, W. Charge Transport in Conduct-\ning Polymers. Faraday Discussions of the Chem. Soc. 1989 ,88,\n223–233.\n[18] Nakata, M.; Taga, M.; Kise, H. Synthesis of Electrical Conduc-\ntive Polypyrrole Films by Interphase Oxidative Polymerization\n– E\u000bects of Polymerization Temperature and Oxidizing Agents.\nPolym. Journ. 1992 ,24, 437–441.\n[19] Meir, Y .; Wingreen, N. S. Landauer Formula for the Current\nThrough an Interacting Electron Region. Phys. Rev. Lett. 1992 ,68, 2512–2515.\n[20] Bao, Z.; Lovinger, A. J.; Dodabalapur, A. Organic Field-E \u000bect\nTransistors with High Mobility Based on Copper Phthalocya-\nnine. Appl. Phys. Lett. 1996 ,69, 3066–3068.\n[21] Parthasarathy, G.; Burrows, P. E.; Khalfin, V .; Kozlov, V . G.;\nForrest, S. R. A Metal-Free Cathode for Organic Semiconduc-\ntor Devices. Appl. Phys. Lett. 1988 ,72, 2138–2140.\n[22] Chen, G.; Shakouri, A. Heat Transfer in Nanostructures for\nSolid-State Energy Conversion. J. of Therm. Transf. 2002 ,124,\n242.\n[23] Chen, G.; Chen, G. Nanoscale Heat Transfer and Information\nTechnology. Appl. Phys. Lett. 2003 ,29, 1–3.\n[24] Haibo, M.; Schollwock, U. Dynamical Simulations of Po-\nlaron Transport in Conjugated Polymers with the Inclusion of\nElectron-Electron Interactions. J. Phys. Chem. A 2009 ,113,\n1360–1367.\n[25] Haug, H. J.; Jauho, A.-P. Quantum Kinetics in Transport and\nOptics of Semiconductors ,2008 , 2nd ed. (Springer, Berlin).\n[26] Giazotto, F.; Heikkilä, T.T.; Luukanen, A.; Savin, A.M.;\nPekola, J. P. Opportunities for Mesoscopics in Thermometry\nand Refrigeration: Physics and Applications. Rev. Mod. Phys.\n2006 ,78, 217–274.\n[27] Dubi, Y .; Di Ventra, M. Colloquium: Heat Flow and Thermo-\nelectricity in Atomic and Molecular Junctions. Rev. Mod. Phys.\n2011 ,83, 131-155.\n[28] Liu, Y . S.; Chen, Y . C. Seebeck Coe \u000ecient of Thermoelectric\nMolecular Junctions: First-Principles Calculations. Phys. Rev.\nB: Condens. Matter Phys. 2009 ,79, 193101.\n[29] Ludovico, M. F.; Lim, J. S.; Moskalets, M.; Arrachea, L.;\nSánchez, D. Dynamical Energy Transfer in ac-Driven Quantum\nSystems. Phys. Rev. B 2014 ,89, 161306.\n[30] Esposito, M. Ochoa, M. A. and Galperin, M., Nature of Ther-\nmal in Strongly Coupled Open Quantum Systems. Phys. Rev.\nB: Condens. Matter Phys. ,2015 ,92, 235440.\n[31] Daré, A.M.; Lombardo, P. Time-Dependent Thermoelectric\nTransport for Nanoscale Thermal Machines. Phys. Rev. B 2016 ,\n93, 035303.\n[32] Arrachea, L.; Bode, N.; von Oppen, F. Vibrational Cooling and\nThermoelectric Response of Nanoelectromechanical Systems.\nPhys. Rev. B 2014 ,90, 125450.\n[33] Zhou, H.; Thingna, J.; Hänggi, P.; Wang, J.-S.; Li, B. Boosting\nThermoelectric E \u000eciency Using Time-Dependent Control. Sci.\nRep.2015 ,5, 14870.\n[34] Segal, D.; Agarwalla, B. K. Vibrational Heat Transport in\nMolecular Junctions. Ann. Rev. Phys. Chem. 2016 ,67, 185–\n209.\n[35] Ye, L.; Zheng, X.; Yan, Y .; Di Ventra, M. Thermodynamic\nMeaning of Local Temperature of Nonequilibrium Open Quan-\ntum Systems Phys. Rev. B 2016 ,94, 245105.\n[36] White, A. J.; Peskin, U.; Galperin, M. Coherence in Charge and\nEnergy Transfer in Molecular Junctions. Phys. Rev. B 2013 ,88,\n205424.\n[37] Eich, F. G.; Di Ventra, M.; Vignale, G. Temperature-Driven\nTransient Charge and Heat Currents in Nanoscale Conductors.\nPhys. Rev. B 2016 ,93, 134309.12\n[38] Wang, R. Q.; Sheng, L.; Shen, R.; Wang, B.; Xing, D. Y . Ther-\nmoelectric E \u000bect in Single-Molecule-Magnet Junctions. Phys.\nRev. Lett. 2010 ,105, 057202.\n[39] Ramos-Andrade, J.P.; Ávalos-Ovando, O.; Orellana, P.A.; Ul-\nloa, S. E. Thermoelectric Transport Through Majorana Bound\nStates and Violation of Wiedemann-Franz Law. Phys. Rev. B\n2016 ,94, 155436.\n[40] Kim, H. S.; Gibbs, Z. M.; Tang, Y .; Wang, H.; Snyder, G. J.\nCharacterization of Lorenz Number with Seebeck Coe \u000ecient\nMeasurement. APL Mater. 2015 ,3, 041506.\n[41] Crépieux, A.; Šimkovic, F.; Cambon, B.; Michelini, F. En-\nhanced Thermopower Under a Time-Dependent Gate V oltage.\nPhys. Rev. B 2011 ,83, 153417.\n[42] Bergfield, J. P.; Sta \u000bord, C. A. Thermoelectric Signatures of\nCoherent Transport in Single-Molecule Heterojunctions. Nano\nLett.2009 ,9, 3072–3076.\n[43] Koole, M.; Thijssen, J. M.; Valkenier, H.; Hummelen, J. C.; Van\nDer Zant, H. S. J. Electric-Field Control of Interfering Trans-\nport Pathways in a Single-Molecule Anthraquinone Transistor.\nNano Lett. 2015 ,15, 5569–5573.\n[44] Valkenier, H.; Guédon, C.M.; Markussen, T.; Thygesen, K.S.;\nvan der Molen, S. J.; Hummelen, J. C. Cross-Conjugation and\nQuantum Interference: A General Correlation? Phys. Chem.\nChem. Phys. 2014 ,16, 653–662.\n[45] Fracasso, D.; Valkenier, H.; Hummelen, J. C.; Solomon, G. C.;\nChiechi, R. C. Evidence for Quantum Interference in Sams of\nArylethynylene Thiolates in Tunneling Junctions with Eutectic\nGa-In (EGaIn) Top-Contacts. J. Amer. Chem. Soc. 2011 ,133,\n9556–9563.\n[46] Aradhya, S. V .; Meisner, J. S.; Krikorian, M.; Ahn,\nS.; Parameswaran, R.; Steigerwald, M. L.; Nuckolls, C.;\nVenkataraman, L. Dissecting Contact Mechanics from Quan-\ntum Interference in Single-Molecule Junctions of Stilbene\nDerivatives. Nano Lett. 2012 ,12, 1643–1647.\n[47] Guédon, C.; Valkenier, H.; Markussen, T.; Thygesen, K.; Hum-\nmelen, J.; van der Molen, S. Observation of Quantum Interfer-\nence in Molecular Charge Transport. Nat. Nano. 2012 ,7, 305–\n309.\n[48] Markussen, T.; Thygesen, K. S. Temperature E \u000bects on Quan-\ntum Interference in Molecular Junctions. Phys. Rev. B 2014 ,89,\n085420.\n[49] Bessis, C.; Rocca, M. L. D.; Barraud, C.; Martin, P.; Lacroix,\nJ. C.; Markussen, T. Probing Electron-Phonon Excitations in\nMolecular Junctions by Quantum Interference. Sci. Rep. 2016 ,\n6, 20899.\n[50] Darwish, N.; Díez-Pérez, I.; Da Silva, P.; Tao, N.; Gooding, J.\nJ.; Paddon-Row, M. N. Observation of Electrochemically Con-\ntrolled Quantum Interference in a Single Anthraquinone-Based\nNorbornylogous Bridge Molecule. Ang. Chem., Int. Ed. 2012 ,\n51, 3203–3206.\n[51] Saygun, T.; Bylin, J.; Hammar, H.; Fransson, J. V oltage-\nInduced Switching Dynamics of a Coupled Spin Pair in a\nMolecular Junction. Nano Lett. 2016 ,16, 2824-2829.\n[52] Díazand, S.; Núñez, Á. S. Current-Induced Exchange Interac-\ntions and E \u000bective Temperature in Localized Moment Systems.\nJ. Phys: Condens. Matt. 2012 ,24, 116001.\n[53] Fransson, J.; Ren, J.; Zhu, J.-X. Electrical and Thermal Control\nof Magnetic Exchange Interactions. Phys. Rev. Lett. 2014 ,113,\n257201.\n[54] Fransson, J.; Zhu, J. X. Spin Dynamics in a Tunnel Junction\nBetween Ferromagnets. New J. Phys. 2008 ,10, 013017.\n[55] Hammar, H.; Fransson, J. Time-Dependent Spin and Transport\nProperties of a Single Molecule Magnet in a Tunnel Junction.\nPhys. Rev. B. 2016 ,94, 054311.[56] Zimbovskaya, N. A. Transport Properties of Molecular Junc-\ntions 2013 , V ol. 254 (Springer, New York).\n[57] Moskalets, M.; Haack, G. Heat and Charge Transport Mea-\nsurements to Access Single-Electron Quantum Characteristics.\nphys. stat. sol. 2016 ,254, 1600616.\n[58] Gaury, B.; Weston, J.; Santin, M.; Houzet, M.; Groth, C.;\nWaintal, X. Numerical Simulations of Time-Resolved Quantum\nElectronics. Phys. Rep. 2014 ,534, 1–37.\n[59] Zimbovskaya, N. A. Seebeck E \u000bect in Molecular Junctions. J.\nPhys: Condens. Matt. 2016 ,28, 183002.\n[60] Simmonds, R. W. Thermal Physics: Quantum Interference\nHeats Up. Nature 2012 ,492, 358–359.\n[61] Heinrich, B. W.; Braun, L.; Pascual, J. I.; Franke, K. J. Protec-\ntion of Excited Spin States by a Superconducting Energy Gap.\nNat. Phys. 2013 ,9, 765–768.\n[62] Heinrich, B. W.; Ahmadi, G.; Müller, V . L.; Braun, L.; Pas-\ncual, J. I.; Franke, K. J. Change of the Magnetic Coupling of a\nMetal-Organic Complex with the Substrate by a Stepwise Lig-\nand Reaction. Nano Lett. 2013 ,13, 4840–4843.\n[63] Heinrich, B. W.; Braun, L.; Pascual, J. I.; Franke, K. J. Tuning\nthe Magnetic Anisotropy of Single Molecules. Nano Lett. 2015 ,\n15, 4024–4028.\n[64] Brumboiu, I. E.; Haldar, S.; Lüder, J.; Eriksson, O.; Herper,\nH. C.; Brena, B.; Sanyal, B. Influence of Electron Correlation\non the Electronic Structure and Magnetism of Transition-Metal\nPhthalocyanines. J. Chem. Theor. Comp. 2016 ,12, 1772–1785.\n[65] Urdampilleta, M.; Klyatskaya, S.; Cleuziou, J.-P.; Ruben, M.;\nWernsdorfer, W. Supramolecular Spin Valves. Nat. Mater. 2011 ,\n10, 502–506.\n[66] Vincent, R.; Klyatskaya, S.; Ruben, M.; Wernsdorfer, W.; Bale-\nstro, F. Electronic Read-Out of a Single Nuclear Spin Using a\nMolecular Spin Transistor. Nature 2012 ,488, 357–360.\n[67] Liu, Z. F.; Wei, S.; Yoon, H.; Adak, O.; Ponce, I.; Jiang, Y .;\nJang, W. D.; Campos, L. M.; Venkataraman, L.; Neaton, J.\nB. Control of Single-Molecule Junction Conductance of Por-\nphyrins via a Transition-Metal Center. Nano Lett. 2014 ,14,\n5365–5370.\n[68] Fransson, J.; Galperin, M. Spin Seebeck Coe \u000ecient of a Molec-\nular Spin Pump. Phys. Chem. Chem. Phys. 2011 ,13, 14350-\n14357.\n[69] Misiorny, M.; Hell, M.; Wegewijs, M. R. Spintronic Magnetic\nAnisotropy. Nat. Phys. 2013 ,9, 801–805.\n[70] Galperin, M.; Nitzan, A.; Ratner, M. A. Heat Conduction in\nMolecular Transport Junctions. Phys. Rev. B. 2007 ,75, 155312.\n[71] Jauho, A. P.; Wingreen, N. S.; Meir, Y . Time-dependent Trans-\nport in Interacting and Noninteracting Resonant-Tunneling Sys-\ntems. Phys. Rev. B. 1994 ,50, 5528–5544.\n[72] Coronado, E.; Day, P. Magnetic Molecular Conductors. Chem.\nRev.2004 ,104, 5419–5448.\n[73] Chen, X.; Fu, Y .-S.; Ji, S.-H.; Zhang, T.; Cheng, P.; Ma,\nX.-C.; Zou, X.-L.; Duan, W.-H.; Jia, J.-F.; Xue, Q.-K. Prob-\ning Superexchange Interaction in Molecular Magnets by Spin-\nFlip Spectroscopy and Microscopy. Phys. Rev. Lett. 2008 ,101,\n197208.\n[74] Chudnovskiy, A. L.; Swiebodzinski, J.; Kamenev, A. Spin-\nTorque Shot Noise in Magnetic Tunnel Junctions. Phys. Rev.\nLett.2008 ,101, 066601.\n[75] Ludwig, T.; Burmistrov, I. S.; Gefen, Y .; Shnirman, A. Strong\nNonequilibrium E \u000bects in Spin-Torque Systems. Phys. Rev. B\n2017 ,95, 075425.\n[76] Hammar, H.; Fransson, J. Transient Spin Dynamics in a Single-\nMolecule Magnet. Phys. Rev. B 2017 ,96, 214401." }, { "title": "1309.2365v1.Spin_rectification_induced_by_dynamical_Hanle_effect.pdf", "content": "arXiv:1309.2365v1 [cond-mat.mes-hall] 10 Sep 2013Spin rectification induced by dynamical Hanle effect\nHiroto Sakimura, Takahiko Matsumoto, and Kazuya Ando∗\nDepartment of Applied Physics and Physico-Informatics,\nKeio University, Yokohama 223-8522, Japan\nAbstract\nDynamic responseof spinaccumulation to a time-dependent m agnetic field has been investigated\nin a ferromagnetic/nonmagnetic bilayer under ferromagnet ic resonance. In this system, magneti-\nzation precession driven by a microwave generates direct-c urrent (dc) and alternate-current (ac)\nspin accumulation in the nonmagnetic layer by the spin pumpi ng. The ac spin accumulation is\ncoupled with the microwave magnetic field through a dynamica l Hanle spin precession, giving rise\nto rectified spin accumulation comparable with the dc spin ac cumulation directly generated by the\nspin pumping.\n∗ando@appi.keio.ac.jp\n1M(t)\nzyxh(t)\nmy(t)δmzδHanle \nFIG. 1: A schematic illustration of the spin pumping and dyna mical Hanle effect. M(t) andh(t)\nare the magnetization and microwave magnetic field, respect ively. The spin pumping creates ac\nspin accumulation δmy(t), which is rectified through the spin precession induced by h(t), giving\nrise to dc spin accumulation δmHanle\nz.\nRectification effects are fundamental in electrical, optical, and mag netic systems. An\nelectrical rectifier is used to convert an alternating current to a d irect current. The\nrectifier essentially strips the high-frequency or alternating part from the incoming cur-\nrent and delivers a low-frequency current, which is based on the tr igonometric relation:\ncosω1tcosω2t= (1/2)[cos(ω1−ω2)t+cos(ω1+ω2)t]; when two waves of frequencies ω1and\nω2are combined, the difference-frequency and sum-frequency ter ms appear. If the frequen-\ncies satisfy ω1=ω2, thetermcos( ω1−ω2)tgives risetoadirect-current (dc) signal. Asimilar\ndynamic response to the product of alternate-current (ac) spin accumulation and a time-\ndependent magnetic field is the origin of a spin rectification effect indu ced by a dynamical\nHanle effect presented in this paper.\nTheHanleeffect refersto thevariationofspin accumulation δminresponse toa magnetic\nfieldHapplied perpendicular to the spin-polarization direction1–3; when a nonmagnet is\nexposed to a magnetic field, spins in the nonmagnet start to preces s around Has∂δm/∂t=\n−γδm×H.\nIn this letter, we show rectification of ac spin accumulation through spin precession in-\nduced by a time-dependent magnetic field: a dynamical Hanle effect. The dynamical Hanle\neffect is discussed in a ferromagnetic/nonmagnetic ( F/N) junction under ferromagnetic res-\nonance, where uniform magnetization precession is driven by a micro wave magnetic field.\nIn this system, the magnetization precession generates dc and ac spin accumulation, or spin\ncurrents, in the Nlayer by the spin pumping.4The dc-component of the spin pumping has\nbeen intensely studied inrecent years.5–11Incontrast, the ac-component ofthe spinpumping\n20200 400 600 800 1000 mzdc \nδ myac \nδ /\n10 -13 10 -12 10 -11 10 -10 10 -9 10 -8 \nτ(s) 10 -3 10 -2 10 -1 0246810 mz (10 3)dc \nδ myac \nδ /\nατ=10 -13 s\nFIG. 2: The ratio |δmy(t)|/δmdc\nzof the magnitude of the ac spin accumulation |δmy(t)|to that\nof the dc spin accumulation δmdc\nzdirectly induced by the spin pumping as a function of the spin\nrelaxation time τin the nonmagnetic layer for 4 πMs= 0.745 T,h= 0.01 mT,α= 0.01, and\nf= 9.4 GHz. The inset shows Gilbert damping constant αdependence of |δmy(t)|/δmdc\nz.\nhas been directly observed only recently.12Here, we show that the ac-component of the spin\npumping not only generates an ac electric voltage through the inver se spin Hall effect but\nalso generates, through the dynamical Hanle spin precession, larg e dc spin accumulation\ncomparable with the dc spin accumulation directly generated by the d c-component of the\nspin pumping. This finding offers a way to generate a large dc electric v oltage from the\nac-component of the spin pumping using the inverse spin Hall effect.\nThe spin pumping in a F/Njunction generates a spin current js(t) and spin accumulation\nδm(t) in the Nlayer.4–11,13–18The spin current density created by the spin pumping is\nexpressed as13,14\njs(t) =/planckover2pi1\n4πg↑↓\neff1\nM2s/parenleftbigg\nM(t)×dM(t)\ndt/parenrightbigg\n, (1)\nwhereg↑↓\neffis the spin pumping conductance and Msis the saturation value of magnetization\nM. Equation (1) shows that the spin pumping generates two types of spin currents; a dc\nspin current with the spin polarization direction along the zaxis, i.e. the magnetization\nprecession axis, and an ac spin current whose spin polarization direc tion oscillates in the\nx-yplane under the ferromagnetic resonance (FMR) condition [see Fig. 1]. In aF/Nthin\nfilm, where the magnetocrystalline anisotropy can be neglected and a microwave magnetic\nfieldh(t) = (hcosωt,0,0) is applied along the xaxis, the dc and ac spin currents with the\nspin polarization direction along zandyaxis at the resonance condition are obtained from\n3Eq. (1) with the Landau-Lifshitz-Gilbert equation by ignoring the se cond-order contribution\nof the precession amplitude as19\njz\ns=g↑↓\neffh2γ2/planckover2pi1/parenleftBig\n2πMsγ+/radicalbig\n4π2Ms2γ2+ω2/parenrightBig\n16πα2/parenleftbig\n4π2Ms2γ2+ω2/parenrightbig , (2)\njy\ns(t) =jy\ns1cosωt−jy\ns2sinωt, (3)\nwhere\njy\ns1=g↑↓\neffhγ/planckover2pi1/parenleftBig\n2πMsγ+/radicalbig\n4π2Ms2γ2+ω2/parenrightBig\n8πα/radicalbig\n4π2Ms2γ2+ω2, (4)\njy\ns2=g↑↓\neffhγ/planckover2pi1ω\n8π/radicalbig\n4π2Ms2γ2+ω2. (5)\nHere,γandαare the gyromagnetic ratio and the Gilbert damping constant, resp ectively. In\ntheNlayer, the dynamics of spin accumulation δm(x,t) = (δmx(t),δmy(t),δmz(t)) induced\nby the spin pumping is obtained from20\n∂δm(x,t)\n∂t=−γδm(x,t)×Heff(t)−δm(x,t)\nτ+D∇2δm(x,t)+js(t)δ(x),(6)\nwhereHeffis the effective magnetic field, τis the spin relaxation time, and Dis the diffusion\ncoefficient in the Nlayer. For simplicity, in the following discussions, we neglect the exter nal\nmagnetic field and consider the system where the Nlayer is thin enough so that the spin\ndiffusion term can be neglected.\nThe magnitude of the ac and dc spin accumulation directly generated by the spin pump-\ning, i.e., thespinaccumulationintheabsenceofthespinprecession: δm×Heff=0, depends\ncritically on the spin relaxation time τin theNlayer. In the absence of the spin precession,\nthe ac spin accumulation δmy(t) induced by the spin pumping is obtained from Eqs. (3) and\n(6) as\nδmy(t) =τ\n1+(ωτ)2[(jy\ns1+ωτjy\ns2)cosωt+(ωτjy\ns1−jy\ns2)sinωt]. (7)\nUsing Eq. (7), we plot the ratio of the ac to dc spin accumulation, |δmy(t)|/δmdc\nz, in Fig. 2,\nwhere the dc spin accumulation δmdc\nzdirectly generated by the spin pumping is obtained\nfrom Eq. (6): dδmzdc/dt= 0 =−δmzdc/τ+jz\ns. Here, 4 πMs= 0.745 T,h= 0.01 mT,\nγ= 1.86×1011T−1s−1,α= 0.01, andf= 9.4 GHz were used for the calculation, where\nω= 2πfandfis the microwave frequency. Figure 2 shows that |δmy(t)|/δmdc\nzchanges\ncritically around τ≃1/ω= 1.7×10−11s. In a material with τ≫1/ω, where the time scale\n410 -13 10 -12 10 -11 10 -10 10 -9 10 -8 \nτ(s) mzdc \nδ mzHanle \nδ (%) /α\n0.001 \n 0.005 0.01 0.02 0.04 0.06 0.08 = 0.1 \n04\n268\nFIG. 3: The ratio δmHanle\nz/δmdc\nzof the dc spin accumulation induced by the Hanle effect δmHanle\nz\nto that induced by the spin pumping δmdc\nzcalculated for different Gilbert damping constants α\nwith 4πMs= 0.745 T,h= 0.01 mT, and f= 9.4 GHz.\nof the oscillation of the ac spin accumulation 1 /ωis much faster than the spin relaxation\ntimeτ, the ac spin accumulation induced by the spin pumping is almost averag ed out,\nwhereas the dc spin accumulation can be accumulated in the time scale ofτ. In contrast,\nin a material with τ≪1/ω, the spins injected into the Nlayer relaxes before cancelling\nout the ac spin accumulation. Here note that the ac spin current jy\nsis much larger than\nthe dc spin current jz\ns, since the cone angle of the magnetization precession is typically\nless than 1◦, resulting the large ac spin accumulation |δmy(t)|compared with the dc spin\naccumulation δmdc\nzin this system. In fact, as shown in the inset to Fig. 2, |δmy(t)|increases\nwith the Gilbert damping constant α, showing that the ac spin accumulation is significant\nat small amplitudes of magnetization precession, since the precess ion amplitude decreases\nwith increasing α.\nThe acspin accumulation δmy(t) induced by the spin pumping creates rectified spin accu-\nmulation δmHanle\nzwith thespin polarizationdirection along the zaxis throughthe dynamical\nHanleeffect; δmy(t)isrectifiedbytheappliedmicrowave magneticfield h(t) = (hcosωt,0,0)\nbecause of the spin-precession term in Eq. (6): −γδm(x,t)×h(t). By neglecting a contri-\nbution from δmzonδmy, thezcomponent of the spin accumulation due to the precession\nofδmy(t) induced by h(t) is obtained from dδmz(t)/dt=γδmy(t)hcosωt−δmz(t)/τ. The\nspin precession term γδmy(t)hcosωtwith the term δmy(t)∝cosωtgives rise to rectified dc\n5010 20 30 40 50 60 70 mzdc \nδ mzHanle \nδ (%) /\n6 4 21.5 1 0.8 0.6 f = 0.45 GHz \n8 10 \n10 -13 10 -12 10 -11 10 -10 10 -9 10 -8 \nτ(s) 2 4 6 8 10 030 60 \nf (GHz) mzdc \nδ mzHanle \nδ (%) /\nFIG. 4: δmHanle\nz/δmdc\nzcalculated for different microwave frequencies fas a function of the spin\nrelaxation time τ. 4πMs= 0.745 T,h= 0.01 mT, and α= 0.05 were used for the calculation. The\ninset shows fdependence of the maximum value of δmHanle\nz/δmdc\nz, i.e.,δmHanle\nz/δmdc\nzatτ= 1/ω.\nspin accumulation δmHanle\nz= (ω/2π)/integraltext2π/ω\n0δmz(t)dt:\nδmHanle\nz=hγτ2(jy\ns1+ωτjy\ns2)\n2/bracketleftbig\n1+(ωτ)2/bracketrightbig. (8)\nIn Fig. 3, we show the ratio of the dc spin accumulation δmHanle\nzinduced by the dynamical\nHanleeffecttothedcspinaccumulation δmdc\nz=τjz\nsdirectlyinducedbythedcspinpumping:\nδmHanle\nz\nδmdc\nz=ατ/parenleftbig\n4π2Ms2γ2+ω2/parenrightbig/bracketleftBig\n2πMsγ+ατω2+/radicalbig\n4π2Ms2γ2+ω2/bracketrightBig\n/bracketleftbig\n1+(ωτ)2/bracketrightbig/bracketleftBig\nω2+2πMsγ/parenleftBig\n2πMsγ+/radicalbig\n4π2Ms2γ2+ω2/parenrightBig/bracketrightBig (9)\nfor 4πMs= 0.745 T and f= 9.4 GHz. Figure 3 shows that δmHanle\nz/δmdc\nzis maximized\naroundτ≃1/ω= 1.7×10−11s. Here note that τ≃1/ωis the spin relaxation time where\nthe ratio of the ac to dc spin accumulation, |δmy(t)|/δmdc\nz, changes drastically [see Fig. 2],\nsuggesting that it is important to consider separately the two situa tions:τ≪1/ωand\nτ≫1/ω. In a material with τ≪1/ω, spins injected into the Nlayer relax before the\nprecession of δmy(t) due toh(t), resulting the suppression of δmHanle\nzas Fig. 3. In contrast,\nin a material with τ≫1/ω,δmHanle\nzis suppressed by the suppression of |δmy(t)|/δmdc\nz\n[see Fig. 2], resulting the peak structure of δmHanle\nz/δmdc\nzwith respect to τ. Therefore,\nthe dynamical Hanle effect is maximized in a material with the spin relaxa tion time of\nτ≃1/ω, showing that the rectification can be controlled by tuning the micro wave frequency\nf=ω/(2π).\n6The rectification effect is also sensitive to the Gilbert damping consta ntαof theF\nlayer as shown in Fig. 3; the dynamical Hanle spin precession is efficient in a system with\nlargeα. This is due to the fact that the ac spin accumulation δmy(t) induced by the spin\npumping becomes significant compared with the dc spin accumulation δmdc\nzwith increasing\nαas shown in the inset to Fig. 2. At first sight, this result seems to imply that the spin\nrectification depends also on the amplitude of the microwave magnet ic fieldh, since the\ncone angle of the magnetization precession decreases with decrea singh, or|δmy(t)|/δmdc\nz\nincreases with decreasing h. However, this compensates the decrease of the spin-precessio n\nterm; by decreasing h, the spin-precession term δm×Heffalso decreases, resulting that the\nrectification of the spin accumulation is independent of has described in Eq. (9).\nThe magnitude of the rectified spin accumulation δmHanle\nzcan be comparable with that\nof the dc spin accumulation δmdc\nzdirectly induced by the dc spin pumping at low microwave\nfrequencies. Figure 4 shows δmHanle\nz/δmdc\nzat different microwave frequencies fcalculated\nusing Eq. (9) with 4 πMs= 0.745 T and α= 0.05. Figure 4 shows that the spin relaxation\ntimeτmaxat which δmHanle\nzis maximized increases with decreasing fbecause of τmax≃1/ω.\nNotably, the maximum value of δmHanle\nz/δmdc\nzincreases with decreasing f[see also the inset\nto Fig. 4], and δmHanle\nzexceeds 50%of δmdc\nzatf= 0.45 GHz. Since the rectification depends\non the magnetic damping αas shown in Fig. 3, δmHanle\nzcan be further enhanced by selecting\na spin injector ferromagnet with a large magnetic damping α, providing a route for realizing\nδmHanle\nz/δmdc\nz>100%.\nIn summary, we have shown that ac spin accumulation generated by the spin pumping\nis rectified through spin precession induced by a microwave magnetic field: the dynamical\nHanle effect. The magnitude of the rectified spin accumulation can be comparable with\nthat of the dc spin accumulation directly induced by the spin pumping. The rectification is\nsensitive to the Gilbert damping constant and microwave frequency , providing a route for\ngenerating giant dc spin accumulation through the dynamical Hanle e ffect.\nThisworkwassupportedbytheCabinetOffice, Government ofJapa nthroughitsFunding\nProgram for Next Generation World-Leading Researchers, the As ahi Glass Foundation, the\nNoguchi Institute, the Murata Science Foundation, and the Mitsu bishi Foundation.\n1F. J. Jedema, A. T. Filip, and B. J. van Wees, Nature 410, 345 (2001).\n72S. P. Dash, S. Sharma, R. S. Patel, M. P. de Jong, and R. Jansen, Nature462, 491 (2009).\n3Y. Fukuma, L. Wang, H. Idzuchi, S. Takahashi, S. Maekawa, and Y. Otani, Nature Mater. 10,\n527 (2011).\n4H. Jiao and G. E. W. Bauer, Phys. Rev. Lett. 110, 217602 (2013).\n5M. V. Costache, M. Sladkov, S. M. Watts, C. H. van der Wal, and B . J. van Wees, Phys. Rev.\nLett.97, 216603 (2006).\n6K. Ando, T. Yoshino, and E. Saitoh, Appl. Phys. Lett. 94, 152509 (2009).\n7K. Ando, J. Ieda, K. Sasage, S. Takahashi, S. Maekawa, and E. S aitoh, Appl. Phys. Lett. 94,\n262505 (2009).\n8O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Ba der, and A. Hoffmann, Phys.\nRev. Lett. 104, 046601 (2010).\n9A. Azevedo, L. H. Vilela-Leao, R. L. Rodriguez-Suarez, A. F. Lacerda Santos, and S. M.\nRezende, Phys. Rev. B 83, 144402 (2011).\n10F. D. Czeschka, L. Dreher, M. S. Brandt, M. Weiler, M. Althamm er, I.-M. Imort, G. Reiss,\nA. Thomas, W. Schoch, W. Limmer, H. Huebl, R. Gross, S. T. B. Go ennenwein, Phys. Rev.\nLett.107, 046601 (2011).\n11E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Appl. Phys. Le tt.88, 182509 (2006).\n12D. Wei, M. Obstbaum, C. Back, and G. Woltersdorf, ArXiv e-pri nts (2013), 1307.2961.\n13Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys. Rev. Let t.88, 117601 (2002).\n14A. Brataas, Y. Tserkovnyak, G. E. W. Bauer, and B. I. Halperin , Phys. Rev. B 66, 060404(R)\n(2002).\n15S. Mizukami, Y. Ando, and T. Miyazaki, Phys. Rev. B 66, 104413 (2002).\n16Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys. Rev. B 66, 224403 (2002).\n17B. Heinrich, Y. Tserkovnyak, G. Woltersdorf, A. Brataas, R. Urban, and G. E. W. Bauer, Phys.\nRev. Lett. 90, 187601 (2003).\n18T. Taniguchi and H. Imamura, Phys. Rev. B 76, 092402 (2007).\n19K. Ando, S. Takahashi, J. Ieda, Y. Kajiwara, H. Nakayama, T. Y oshino, K. Harii, Y. Fujikawa,\nM. Matsuo, S. Maekawa, E. Saitoh, J. Appl. Phys. 109, 103913 (2011).\n20K. Ando, S. Takahashi, J. Ieda, H. Kurebayashi, T. Trypiniot is, C. H. W. Barnes, S. Maekawa,\nand E. Saitoh, Nature Mater. 10, 655 (2011).\n8" }, { "title": "2009.04423v4.Superconductivity_enhanced_spin_pumping__Role_of_Andreev_resonances.pdf", "content": "Superconductivity-enhanced spin pumping: Role of Andreev resonances\nMostafa Tanhayi Ahari and Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe describe a simple hybrid superconductor jferromagnetic-insulator structure manifesting spin-\nresolved Andreev bound states in which dynamic magnetization is employed to probe spin related\nphysics. We show that, at low bias and below Tc, the transfer of spin angular momentum pumped\nby an externally driven ferromagnetic insulator is greatly a\u000bected by the formation of spin-resolved\nAndreev bound states. Our results indicate that these bound states capture the essential physics\nof condensate-facilitated spin \row. For \fnite thicknesses of the superconducting layer, comparable\nto the coherence length, resonant Andreev bound states render highly transmitting subgap spin\ntransport channels. We point out that the resonant enhancement of the subgap transport channels\nestablishes a prototype Fabry-P\u0013 erot resonator for spin pumping.\nIntroduction .|Spatial variations in the superconduct-\ning order in a \fnite region lead to the formation of spin-\ndegenerate Andreev bound states (ABSs) with discrete\nexcitation energies below the superconducting gap [1].\nAn externally applied magnetic \feld or proximity to a\nferromagnetic order, on the other hand, can induce spin\nsplitting in the ABSs that results into spin-resolved ABSs\n[2]. In this paper, we consider a normal metal (N) sand-\nwiched between a superconductor (S) and a ferromag-\nnetic insulator (FI) that serves as a simple platform with\nspin-resolved ABSs, which are localized in the N layer.\nThe nonequilibrium pure spin current engendered from\nthe externally driven FI|spin pumping|is utilized to\nprobe spin transport in an S jNjFI hybrid structure. The\nspin pumping generated from a time-dependent magne-\ntization, on the other hand, is a \row of spin angular\nmomentum into adjacent materials that dissipates en-\nergy of the ferromagnet [3]. We suggest that the mag-\nnetic damping increase in superconducting hybrid multi-\nlayers NbjNi80Fe20and NbNjGdN reported, respectively,\nin Refs. [4] and [5] may be attributed to the resonant\nenhancement of the spin pumping discussed here.\nIn the context of superconducting spintronics, combin-\ning ans-wave superconducting order, favoring electrons\nto form a singlet state, with a ferromagnetic order, fa-\nvoring spin alignment, leads to a powerful enhancement\nor reduction of angular momentum transfer [6, 7]. The\nangular momentum transfer, as a central e\u000bect in spin-\ntronics, is greatly modi\fed on account of two major un-\nderlying causes: the itinerant spin-polarized quasiparti-\ncles (QPs) with long spin-coherence lengths [8, 9] and\nthe creation of spin-triplet Cooper pairs [10{12] induced\nat highly spin-active regions or complex magnetic mul-\ntilayers [13]. Here, the spin pumping in an S jNjFI hy-\nbrid structure, however, is an interplay between spin-\npolarized QPs and spin-triplet Cooper pairs, which are\ndynamically generated by the excited FI [14, 15]. The\nsubgap ABSs accommodate the spin-polarized QPs and\nspin-triplet Cooper pairs that for a su\u000eciently thin S\nlayer can tunnel across and contribute to the spin current.\nTo collect the spin current we have placed a spin reser-\nvoir N r, comprising an N rjSjNjFI structure (see Fig. 1).\nFIG. 1. Schematic sketch for an N rjSjNjFI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum, Is, that is carried by the QPs into the\nreservoir N r(spin pumping). For \fnite thicknesses of the S\nand N layers ( dS;dN\u0018coherence length) the resonant ABSs\nprovide highly transmitting transport channels for spin pump-\ning. Here, for example, we show four resonant channels with\ntwo positive-energy ABSs. The energies are measured with\nrespect to the Fermi energy EFdepicted in the middle.\nWhile spin pumping has been considered as a new probe\nof spin dynamics in a superconducting thin \flm [16] and\ntunable pure spin supercurrents [17], here we propose to\nstudy spin-resolved Andreev resonances by means of spin\npumping.\nSpin-resolved ABS .|We will determine the subgap\nspin-resolved ABSs by studying the resonant conditions\nfor the spin transport in an N rjSjNjFI hybrid struc-\nture. In order to establish the pumped spin current in\nthe presence of superconducting and dynamic ferromag-\nnetic orders, we solve a time-dependent scattering prob-\nlem [15, 18], which accounts for the relevant processes\nsuch as Andreev re\rection (AR), ABS, and a dynamic\ntriplet-paring generation. For subgap energies, the full\nscattering matrix develops peak structures marking res-\nonant bound states [19] (ABSs), which result in highly\ntransmitting subgap transport channels for the spin \row.\nWhile we identify the Andreev resonances in the ballistic\nregime, we do not expect their perturbation by a weak\ndisorder to signi\fcantly modify the spectroscopic aspects\nof spin pumping (which should be governed by the un-\nderlying dynamic mixing between the superconductingarXiv:2009.04423v4 [cond-mat.supr-con] 19 Mar 20212\nFIG. 2. Pro\fle of Re g\"#fordN=\u00180= 1. (a) Assuming \fxed\nS layer thickness ( dS=\u00180= 1:5), the spectral overlap of the\nspin-split Andreev levels can be modi\fed by \u0001 ex. (b) At the\n\fnite thicknesses of the S, dS&\u00180, the Andreev resonances\nestablish highly transmitting transport channels. Here, we\nhave set \u0001 ex=EF= 0:4.\nsinglet and triplet components). To capture the essen-\ntial e\u000bect of interfacial scattering on the spin-pumping\nenhancement, we will consider an interface disorder at\nx=dS.\nWe proceed with establishing notations and key fea-\ntures of the scattering. The incident QPs from the reser-\nvoir onto the N rjS interface can either transmit across\nthe S or re\rect as holes back into the reservoir, a process\nknown as AR. The AR amplitude for an incident QP with\nenergy\"is given by\nr1\nA=(\ne\u0000iarccos (\"=\u0001);j\"j<\u0001;\n\"\u0000\"p\n1\u0000\u00012=\"2\n\u0001;j\"j>\u0001;(1)\nwhere \u0001 is the superconducting pair-potential [2]. The\nAR amplitude given in Eq. (1), by focusing on the inter-\nface, assumes a bulk S (a thick S layer) [15]. A nonzero\nprobability of QP transmission for the thin S as well as\nan interfacial disorder can reduce the AR probability,\nwhich we address below. The transmitted QPs with spin\n\u001b2f\";#gpropagate through the normal metal junction\nand acquire a spin-dependent phase ei#\u001bupon re\rection\nfrom the FI interface, where the FI region is considered\nan exchange-splitting insulator for the QPs [21]. This\nrenders the full scattering matrix to be merely a re\rec-\ntion matrix, which is block diagonal in the spin space\ndue to the conservation of the QP spin during each indi-\nvidual scattering event. Consequently, multiple ARs at\nthe NjS along with spin-dependent re\rections at the N jFI\ninterfaces constitute the full scattering matrix.\nThe spin-active interface N jFI, upon re\rection, rotates\na noncollinear QP spin around the FI magnetization axis,\nwhich in turn for a driven magnetization leads to genera-\ntion of a nonequilibrium spin current detected in N r[22].\nThe conductance determining the transport of this spin\ncurrent, known as the mixing conductance g\"#, is givenby\ng\"#=X\nn;m\u0000\n\u000em;n\u0000r\"\nee;mnr#\u0003\nee;mn +r\"\nhe;mnr#\u0003\nhe;mn\u0001\n;(2)\nwherer\u001b\neeandr\u001b\nherepresent the total spin- \u001belectron-to-\nelectron and electron-to-hole re\rection matrices in the\nSjNjFI hybrid structure, respectively [14, 18]. The in-\ndicesnandmrefer to transport channels the number\nof which can be determined by transverse thickness of\nthe normal metal layer [2]. The re\rection matrices in\nEq. (2) are written in the basis where the spin quan-\ntization axis is parallel to the magnetization in the FI\n(the exact expressions for the re\rection amplitudes in\nthe ballistic regime can be found in the Supplemental\nMaterial [15]). In the following, we provide the results\nfor the single-channel scattering and postpone discussing\nthe case of multichannel scattering to the Supplemental\nMaterial [15].\nGenerically, the mixing conductance given in Eq. (2) is\na complex number the real part of which governs the av-\nerage spin pumping and the associated Gilbert damping\n[18]. After a straightforward calculation in the ballistic\nregime, three distinct regimes are recognized for subgap\nenergies:\nRe g\"#\u00198\n><\n>:1\u0000cos# dS\u001c\u00180\n\r2=(\r2+\f2)dS\u0018\u00180\n0 dS\u001d\u00180; (3)\nwhere\r = exp[\u00002p\n1\u0000\"2=\u00012\n0dS=\u00180],\f =\u0002\ncos (4dN\"=\u00180\u00010+ 2\u0012A)\u0000cos#\u0003\n=4 sin#sin2\u0012A, and\n\u0012A\u0011 \u0000 arccos (\"=\u00010). The mixing angle de\fned as\n#\u0011#\"\u0000##is controlled by the FI exchange interaction\n\u0001ex[15]. Here, \u0001 0and\u00180are the superconducting gap\nand coherence length at zero temperature, respectively.\nEvidently, Re g\"#displays a resonant behavior associated\nwith the intermediate S thickness. The resonant energy\nlevels correspond to the ABS energies \"Adetermined by\n(\f= 0):\n\"A= \u0001 0signh\nsin\u00122dN\"A\n\u00180\u00010\u0006#\n2\u0013i\ncos\u00122dN\"A\n\u00180\u00010\u0006#\n2\u0013\n:\n(4)\nWhen the exchange interaction is absent, that is, #= 0,\nEq. (4) yields a pair of solutions ( \u0000\"A;\"A), which is a\nconsequence of the particle-hole symmetry imposed on\nthe scattering formalism [15]. A nonzero mixing angle\n#, on the other hand, by lifting the spin degeneracy of\neach level results in the spin-resolved ABSs with the fol-\nlowing energies (\u0000\"\u0006\nA;\"\u0006\nA). As an outcome, the spectral\noverlap of the spin-split bound states can be controlled\nby the exchange interaction of the FI [e.g., see Fig. 2(a)].\nWe emphasize that, following Eq. (3), only at dS\u0018\u00180\nthe resonant ABSs establish highly transmitting trans-\nport channels for the spin \row, which is greatly enhanced3\nFIG. 3. Geometric resonances of Re g\"#. The Fabry-P\u0013 erot\noscillations happen when the corresponding AR probability\njrAjis nonzero (the inset plot shows the probability of AR\nvs energy). For a \fxed \u0001 ex, the Fabry-P\u0013 erot oscillations are\ndetermined by dNthat for\" > \u00010get modulated with a\nfrequency determined by dS(i.e., energies for which jrAj= 0).\nHere, we have adopted the following parameters: dS=\u00180= 1,\ndN=\u00180= 10, and \u0001 ex=EF= 0:2.\ncompared to either a bulk or no S layer [see Fig. 2(b)].\nThis is one of the main results of this paper.\nIn the limit where dN!0, Eq. (4) can be reduced\nto the well known result [23] for ABSs with magnetically\nactive interfaces, \"A=\u00010=\u0006cos#=2. We highlight the\nfact that Eq. (4) is a condition for a constructive quan-\ntum interference in a Rowell{McMillan process for the\nQPs [12], that is, four times crossing Nwith two An-\ndreev conversions as well as two re\rections from FI, once\nas electron and once as hole. It is clear that a construc-\ntive interference for QPs inside the normal layer remains\nintact as long as the probability of the AR is nonzero.\nThis captures the essential physics of a two-mirror Fabry-\nP\u0013 erot resonator with a resonator length 2 dN, which we\nshall describe now. The N jFI and SjN interfaces operate\nas \\mirrors\" for the QPs that give rise, respectively, to\na spin-dependent specular re\rection (a spin-dependent\nmirror) and a phase-conjugating mirror, which retrore-\n\rects electrons with energy EF+\"Aas holes with energy\nEF\u0000\"A[24]. The resonant enhancement of a Fabry-\nP\u0013 erot device occurs when its mirrors have a near unity\nre\rection probability [25]. Here, the N jFI interface re-\n\rects all the incident QPs with probability 1, while the\nretrore\rection probability of the S jN interface, on the\nother hand, is determined by the AR probability. There-\nfore, the Fabry-P\u0013 erot enhancement is in accordance with\nthe AR amplitude, which for an S with a thickness dS\n[15] is given by\nrA=(1\u0000\r)r1\nA\n1\u0000\r(r1\nA)2: (5)\nFor subgap energies, in the limiting case of dS> \u00180, we\nget\r\u001c1 or equivalentlyjrAj\u00191, for which the resonant\nFIG. 4. The normalized e\u000bective conductance, ~ g\"#\ne\u000b, for sub-\ngap temperatures and intermediate thickness dS\u0018\u00180showing\na signi\fcant enhancement relative to the normal case near Tc.\nAs the S gap shrinks, with the increase of temperature, the\nenhancement peak location creeps up to higher dS, which ul-\ntimately diminishes for higher temperatures near Tc. Here,\nwe have set dN=\u00180= 2:5. The inset shows that the enhance-\nment feature happens for all 0 \u0014dN=\u00180\u00145 whendS=\u00180= 1.\nFor the above plots, we have used \u0001 ex=EF= 0:2.\nenhancement of a Fabry-P\u0013 erot device is expected. For en-\nergies above the gap, on the other hand, the AR re\rection\namplitude decreases. This results in highly transmitive\nchannels near \"&\u00010, which rapidly decline for higher\nenergies (see Fig. 3).\nIn addition to the QP interference in the normal layer\n(Rowell{McMillan resonance), which leads to the mix-\ning conductance oscillation, above the gap a quantum\ninterference inside the S layer can take place. An in-\ncident electronlike QP interferes with a holelike QP re-\n\rected from the other S jN interface, which may result in\njrAj= 0. The process known as Tomasch oscillations [26]\noccurs for QP energies \"n=\u00010=p\n1 + (n\u0019\u00180=2dS)2with\nintegern, which modulates the amplitude for the Fabry-\nP\u0013 erot oscillations. The Rowell{McMillan and Tomasch\nbased geometric resonances of Re g\"#are shown in Fig. 3.\nSpin pumping enhancement .|The spin pumping is\ngenerated by the variations in the magnetization direc-\ntionm(t) [3, 10, 27]. For su\u000eciently slow variations,\nto the \frst order in the pumping parameter frequency\nj@tmj, the spin pumping can be written in terms of the\ninstantaneous mixing conductance in the magnetization\ncoordinate system, that is, Eq. (2). Consequently, the\nspin-pumping current, assuming no voltage bias [18], is\ngiven by\nIs(t) =1\n4\u0019g\"#\ne\u000bm\u0002@tm; (6)\nwhere the e\u000bective mixing conductance is de\fned as fol-4\nlows:\ng\"#\ne\u000b\u0011Z1\n\u00001d\"@\"f(\") Re g\"#: (7)\nHere,f(\") = (1 +e\"=kBT)\u00001is the Fermi-Dirac distri-\nbution and in order to properly take into account the\ntemperature dependence of the S order, we have consid-\nered a temperature{dependent S gap \u0001( T) [21].\nIn accordance with Eq. (6), the spin{pumping current\nin the direction of m\u0002@tmis simply determined by g\"#\ne\u000b,\nwhich can be regarded as the mixing conductance in the\ntemperature domain. In order to focus on the role of ABS\nresonances, we have de\fned normalized e\u000bective conduc-\ntance ~ g\"#\ne\u000b\u0011g\"#\ne\u000b(dS)=g\"#\ne\u000b(dS= 0). Here, g\"#\ne\u000b(0) yields\nthe e\u000bective conductance in the normal state. The re-\nsultant normalized conductance is plotted in Fig. 4. We\n\fnd that, for subgap temperatures, ~ g\"#\ne\u000bshows a signi\f-\ncant enhancement at the \fnite thickness of the S layer,\ni.e.,dS\u0018\u00180. The enhancement is optimal at the midgap\ntemperatures and diminishes down to unity (the normal\ncase) upon approaching Tc.\nBefore closing this paper, we explore the e\u000bect of a bar-\nrier potential at the S jN interface. Physically, this can\noriginate from a thin oxide layer or a localized disorder\non the interface. The essential e\u000bects of the interfacial\nscattering, caused by this layer, can be captured by a\npotential of the form Z~vF\u000e(x\u0000dS), where the dimen-\nsionless parameter Zdetermines strength of the barrier.\nHere,vFis the Fermi velocity and \u000e(x) is the Dirac delta\nfunction.\nAt the SjN interface, a nonzero Zresults in reduced\ntransmission and AR probabilities by introducing ordi-\nnary electron and hole re\rections [3]. This partially re-\n\rective interface can lead to the formation of normal\nbound states localized in the N layer. In contrast, the\ncase of zero barrier ( Z= 0) leads to ABSs only. For sub-\ngap energies, with the increase of Zthe ordinary re\rec-\ntion probability surpasses the AR probability [3], which\nleads to ABSs to be pushed away from zero (towards the\ncontinuum of states above the gap) and replaced with\nordinary bound states [15]. Consequently, in the strong\nbarrier limit Z > 1, the subgap transport channels are\ndue to the normal bound{state resonances, where super-\nconductivity suppresses the spin pumping (see Fig. 5).\nWe point out that the e\u000bective conductance is normal-\nized with respect to g\"#\ne\u000b(dS= 0), which accounts for the\ncontribution of the normal bound{state resonances. The\nweak barrier limit ( Z < 1), on the other hand, can ef-\nfectively be described within the zero{barrier limit by\nan increased dNand \u0001 ex(due to the multiple ordinary\nre\rections) [see Fig. 5 (inset)].\nConclusion and discussion .|We have shown that su-\nperconductivity can greatly a\u000bect spin pumping due to\nthe formation of the resonant ABSs, which result in\nhighly transmitting spin transport channels when dS\u0018\nFIG. 5. The normalized e\u000bective conductance is shown for\nvarious barrier potential strengths Z, which are plotted by\nincreasingZin increments of 0 :5 from 0 to 2 :5. The enhance-\nment feature ceases to exist for strong barriers Z > 1. The\ninset plot shows that the simultaneous increase of dNand\n\u0001exwithZ= 0 can account for the conductance in the weak\nbarrier case Z <1. The red plots are generated by simultane-\nously increasing both the exchange interaction \u0001 ex=EFfrom\n0:1 to 0:6 in increments of 0 :1 anddN=\u00180from 3 to 4 :5 in\nincrements of 0 :3. In the above plots, we have set dN=\u00180= 3,\n\u0001ex=EF= 0:1.\n\u00180. This can be manifested experimentally by an in-\ncrease in Gilbert damping of the FI dynamics. The\nGilbert damping enhancement in an NbN jGdN struc-\nture has been observed to peak for subgab temperatures\n[5], where NbN is an s{wave superconductor with coher-\nence length of\u00185 nm that is adjacent to a ferromag-\nnetic{insulator \flm, GdN. The enhancement takes place\nfordS= 10 nm and it is suppressed for dS= 2 nm. We\nbelieve that this is in good agreement with our results,\ne.g., Fig. 4. On the other hand, unfolding N rjSjNjFI e\u000bec-\ntively maps our setup to an N rjSjFjSjNrstructure, where\nF stands for a normal metal with a ferromagnetic order.\nThe hybrid structure N rjNbjNi80Fe20jNbjNrstudied in\nRef. [4] shows a signi\fcant spin{pumping enhancement\nonly when the Nb thickness is roughly equal to its coher-\nence length ( dS\u0019\u00180= 30 nm). In order to reveal a large\nenhancement, they have utilized a range of spin{sink ma-\nterials N rwith spin{orbit interaction, such as Pt, W, or\nTa. In Ref. [17] the same hybrid structure of Ref. [4] has\nbeen used except for a particularly magnetized spin sink\nPt/Co/Pt. It is shown that the spin{pumping e\u000eciency\nacross Nb is tunable by controlling the magnetization di-\nrection of Co.\nFurthermore, hybrid Josephson junctions realizing\nABSs with near unity transmission probability for charge\ntransport have been proposed to coherently manipu-\nlate quantum{information devices such as Andreev{level\nqubits [29, 30]. From this standpoint, unfolding our setup\nrealizes a magnetically active Josephson junction [8] with\nresonant transport channels, which in turn can provide a\nspintronic paradigm for a coherent manipulation of quan-5\ntum{information devices involving ABSs. Note added :\nRecently, we became aware of an interesting and closely\nrelated work by Silaev [31].\nIt is a pleasure to acknowledge discussions with C. Ci-\nccarelli and W. A. Robinson who drew our attention to\nthis problem. MTA wishes to thank S. Tanhayi Ahari for\nuseful comments on the paper. This work is supported\nby the U.S. Department of Energy, O\u000ece of Basic Energy\nSciences under Grant No. DE{SC0012190.\n[1] J.-D. Pillet, C. H. L. Quay, P. Mor\fn, C. Bena, A. Levy\nYeyati, and P. Joyez, Nat. Phys. 6, 965 (2010); J. Schin-\ndele, A. Baumgartner, R. Maurand, M. Weiss, and C.\nSch onenberger, Phys. Rev. B 89, 045422 (2014); J. A.\nSauls, Phil. Trans. R. Soc. A. 37620180140 (2018)\n[2] E. J. H. Lee, X. Jiang, M. Houzet, R. Aguado, C. M.\nLieber, and S. De Franceschi, Nature Nanotechnology, 9,\n79 (2014); F. Pawlicki and I. Weymann, Phys. Rev. B 98,\n085411 (2018)\n[3] Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, and\nBertrand I. Halperin, Rev. Mod. Phys. 77, 1375 (2005)\n[4] K-R Jeon, C. Ciccarelli, A. J. Ferguson, H. Kurebayashi,\nL. F. Cohen, X. Montiel, M. Eschrig, J. W. A. Robinson,\nand M. G. Blamire, Nat. Mater. 17, 499{503 (2018)\n[5] Y. Yao, Q. Song, Y. Takamura, and et al, Phys. Rev. B\n97, 224414 (2018)\n[6] S. Kashiwaya, Y. Tanaka, N. Yoshida, and M. R. Beasley,\nPhys. Rev. B 60, 3572 (1999)\n[7] J. Linder, J. W. A. Robinson, Nature Physics, 11, 307{315\n(2015)\n[8] M. Fogelstr om, Phys. Rev. B 62, 11812, (2000)\n[9] M. Eschrig, J. Kopu, J. C. Cuevas, G. Sch on, Phys. Rev.\nLett. 90, 137003 (2003); G. Metalidis, M. Eschrig, R.\nGrein, G. Sch on, Phys. Rev. B 82, 180503(R) (2010); H.\nYang, S-H Yang, S. Takahashi, S. Maekawa, and S. S. P.\nParkin , Nat. Mater. 9, 586 (2010)\n[10] A. Brataas and Y. Tserkovnyak, Phys. Rev. Lett. 93,\n087201 (2004)\n[11] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys.\nRev. Lett. 86, 4096 (2001)\nM. Houzet and A. I. Buzdin, Phys. Rev. B 76, 060504(R)\n(2007)\n[12] M. Eschrig, Phil. Trans. R. Soc. A 376: 20150149 (2018)\n[13] T. Vezin, C. Shen, J. E. Han, and I. \u0014Zuti\u0013 c, Phys. Rev. B\n101, 014515 (2020)\n[14] T. Yokoyama, and Y. Tserkovnyak, Phys. Rev. B 80,\n035408 (2009)\n[15] See the Supplemental Material at\nhttps://link.aps.org/supplemental/DOI for a discus-\nsion of multichannel scattering, exact expressions of the\nre\rection amplitudes, and a heuristic argument for the\ndynamic generation of triplet pairing. We also provide a\nbrief discussion of the ABS modi\fcation in the presence\nof the interface barrier potential.\n[16] M. Inoue, M. Ichioka, and H. Adachi, Phys. Rev. B 96,\n024414 (2017)\n[17] K-R Jeon, X. Montiel, S. Komori, C. Ciccarelli, J. Haigh,\nH. Kurebayashi, L. F. Cohen, A. K. Chan, K. D. Stenning,\nC-M Lee, M. G. Blamire, and J. W. A. Robinson, Phys.Rev. X 10, 031020 (2020)\n[18] H. J. Skadsem, A. Brataas, J. Martinek, and Y.\nTserkovnyak, Phys. Rev. B 84, 104420 (2011)\n[19] B. Belchev, S.G. Neale, M.A. Walton, Can. J. Phys. 89,\n11 (2011)\n[20] Y. Nazarov, Y. Blanter. Quantum Transport: Introduc-\ntion to Nanoscience. Cambridge: Cambridge University\nPress (2009)\n[21] We consider an energy gap of 1 :1EFfor up-spin and\n1:1EF+ 2\u0001 exfor down-spin electrons. The Fermi en-\nergyEFis assumed to be identical for the di\u000berent lay-\ners,EF= 5 eV. Furthermore, we have assumed Nb as\nthe S with Tc= 9 K,\u00180= 30 nm, and \u0001 0= 3 meV.\nWe have adopted the following empirical parametrization\nfor a temperature-dependent superconductor gap for Nb,\n\u0001(T) = \u0001 0(1\u0000T=Tc )1:94(1 + 2:17T=Tc ), that is taken\nfrom R. C. Dougherty, J. D. Kimel. Superconductivity Re-\nvisited. (CRC Press, Boca Raton, FL, 2012)\n[22] A. Brataas, Y. V. Nazarov, and G. E. W. Bauer, Phys.\nRev. Lett. 84, 2481 (2000)\n[23] F. H ubler, M. J. Wolf, T. Scherer, D. Wang, D. Beck-\nmann, and H. v. L ohneysen, Phys. Rev. Lett. 109, 087004\n(2012)\n[24] H. van Houten, C. W. J. Beenakker, Physica B: Con-\ndensed Matter 175, Issues 1{3, (1991)\n[25] N. Ismail, C. C. Kores, D. Geskus, and M. Pollnau, Op-\ntics Express Vol. 24, Issue 15, (2016)\n[26] W. J. Tomasch, Phys. Rev. Lett. 16, 16 (1966).\n[27] K. Xia, P. J. Kelly, G. E. W. Bauer, and I. Turek, Phys.\nRev. Lett. 89, 166603 (2002)\n[28] G. E. Blonder, M. Tinkham, and T. M. Klapwijk. Phys.\nRev. B 25, 4515 (1982)\n[29] F. Nichele, E. Portoles, A. Fornieri, A. M. Whiticar, A.\nC. C. Drachmann, S. Gronin, T. Wang, G. C. Gardner, C.\nThomas, A. T. Hatke, M. J. Manfra, and C. M. Marcus,\nPhys. Rev. Lett. 124, 226801 (2020)\n[30] J. Michelsen, V. S. Shumeiko, and G. Wendin Phys. Rev.\nB77, 184506 (2008); C. Janvier, L. Tosi, L. Bretheau, C.\nO. Girit, M. Stern, P. Bertet, P. Joyez, D. Vion, D. Esteve,\nM. F. Go\u000bman, H. Pothier, C. Urbina, Science 349, Issue\n6253, (2015)\n[31] M. A. Silaev, Phys. Rev. B. 102, 180502(R) (2020)1\nSupplementary Material for\nSuperconductivity-enhanced spin pumping: Role of Andreev resonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nIn this supplementary material, we discuss a QP scattering formalism that takes into account the dynamic mag-\nnetization of the FI and the s-wave order of the S in an N rjSjNjFI hybrid structure, see Fig. S1. We will derive\nanalytical expressions for the scattering amplitudes, ABS modi\fcation under an interfacial disorder, and elaborate on\nthe dynamic generation of the triplet Cooper pairs in our setup.\nSpin-dependnet scattering .|In the context of QP scattering in hybrid structures generic boundary conditions perti-\nnent to interface scattering and disorder can be taken into account by insertion of a perfect spacer N between adjacent\nmaterials [1, 2]. As such, the full scattering process can be broken down into studying N rjSjN, and NjFI structures\nseparately. We bear in mind that propagation through the N layer mixes neither particle-hole nor spin degrees of\nfreedom. For the N rjSjN structure, following BTK [3], we assume a uniform superconducting pair-potential \u0001 inside\nthe S layer that vanishes in the N layer. This way the scattering process takes place only at the interfaces, which\nessentially contains QPs propagating into the S and Andreev retrore\rection. In a more realistic (dirty) N jS interface,\nhowever, there is a possibility of a specular re\rection for electrons and holes. In order to model the partially re\rective\ninterface, we can consider a barrier potential in the form of V=Z1~vF\u000e(x) andZ2~vF\u000e(x\u0000dS). A \fniteZ1, leads to\nquantum interference with the Fermi wavelength and does not a\u000bect ABS. In the following, we set Z1= 0 and focus\non the e\u000bect of a nonzero Z2\u0011Z.\nInitially, we study point contact scattering in a 1D wire. Physically, this can be justi\fed by imposing a constriction,\nfor which the transverse thickness of the N layer is on the order of the Fermi wavelength [1]. Later on, we will relax\nthis condition by considering a 3D geometry for which the incident QPs can have a transverse momentum.\nWe begin with the BdG formalism that accounts for symmetries of the superconducting part for the electron and\nhole wave functions,\nHBdG\t =\"\t;withHBdG=\u0012^H0^\u0001\n^\u0001y\u0000^H\u0003\n0\u0013\n; (S1)\nwhere ^H0=\u0000\np2=2m+V\u0000EF\u0001\nis the single-electron Hamiltonian matrix and ^\u0001 = \u0001i\u001by. Here, we assume a real-\nvalued pairing potential \u0001. The Pauli matrix \u001byis acting on the spin space f\";#gof conduction electrons and holes.\nAbove eigenvalue equation admits the following basic electron and hole QP modes with spin \u001b:\n\t\u0006\nn;e\u001b=\u0012\n\u001f\u001b\nA(\")(\u0000i\u001by)\u001f\u001b\u0013\ne\u0006iqexeikk\u0001rk;\n\t\u0006\nn;h\u001b=\u0012A\u0003(\u0000\")(\u0000i\u001by)\u001f\u001b\n\u001f\u001b\u0013\ne\u0006iqhxeikk\u0001rk; (S2)\nwhere\u001f\"=\u00001\n0\u0001\n,\u001f#= (\u0000i\u001by)\u001f\", are two dimensional spinors, and we have\n~2\n2mq2\ne;h=EF\u0000En+\u0010e;h\n;\n\u0011(\nip\n\u00012\u0000\"2; ifj\"j<\u0001\n\"p\n1\u0000\u00012=\"2;ifj\"j>\u0001; A (\") =(\ne\u0000iarccos (\"=\u0001);j\"j<\u0001\n\"\u0000\"p\n1\u0000\u00012=\"2\n\u0001;j\"j>\u0001; (S3)\nwhereEn\u0011~2k2\nk=2m, and\u0010e;h=\u00061. In the normal layer, the wave number is given by ke;h=qe;h(\u0001!0). The\nconservation of the QP spin leads to a block diagonal full scattering matrix in the spin space. Whence, we can work\nwith the following 2 dimensional eigen vectors,\n\t\u0006\nSne=\u00121\nA(\")\u0013\ne\u0006iqexeikk\u0001rk;\n\t\u0006\nSnh=\u0012A(\")\n1\u0013\ne\u0006iqhxeikk\u0001rk: (S4)\nIt is clear that outside of the S, where \u0001 !0, we getA(\")!0, leading to normal region's basis elements,\n\t\u0006\nNne=\u00121\n0\u0013\ne\u0006ikex\bn(y;z);\n\t\u0006\nNnh=\u00120\n1\u0013\ne\u0006ikhx\bn(y;z); (S5)2\n𝑐\"#\n𝑐\"$\n𝑐%$\n𝑐%#𝑐\"#\n𝑐\"$\n𝑐%$\n𝑐%#\n𝑍' 𝑍(Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥\ng\"#⇤\n. (1)\nThe quantity x, y, z Re⇥\ng\"#⇤\nrefers to the real part of the\nspin-mixing conductance g\"#=T r⇥\n1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤\n,\nwhere r\u0000\neeand r\u0000\u0000\u0000\nhe(\u00002{\",#}) represent the electron-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥g\"#⇤. (1)\nThe quantity x, y, z Re⇥g\"#⇤refers to the real part of the\nspin-mixing conductance g\"#=T r⇥1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤,\nwhere r\u0000\neeand r\u0000\u0000\u0000\nhe (\u00002{\",#}) represent the electron-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥\ng\"#⇤\n. (1)\nThe quantity Re⇥\ng\"#⇤\nrefers to the real part of the spin-\nmixing conductance g\"#=T r⇥\n1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤\n,w h e r e\nr\u0000\neeand r\u0000\u0000\u0000\nhe(\u00002{\",#}) represent the electron-to-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥g\"#⇤. (1)\nThe quantity Re⇥g\"#⇤refers to the real part of the spin-\nmixing conductance g\"#=T r⇥1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤,w h e r e\nr\u0000\neeand r\u0000\u0000\u0000\nhe (\u00002{\",#}) represent the electron-to-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xzy\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥g\"#⇤. (1)\nThe quantity Re⇥g\"#⇤refers to the real part of the spin-\nmixing conductance g\"#=T r⇥1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤,w h e r e\nr\u0000\neeand r\u0000\u0000\u0000\nhe (\u00002{\",#}) represent the electron-to-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥g\"#⇤. (1)\nThe quantity Re⇥g\"#⇤refers to the real part of the spin-\nmixing conductance g\"#=T r⇥1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤,w h e r e\nr\u0000\neeand r\u0000\u0000\u0000\nhe (\u00002{\",#}) represent the electron-to-Superconductivity-enhanced spin pumping: The role of Andreev bound-state\nresonances\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nWe study spin pumping in a hybrid superconductor |ferromagnetic-insulator structure. We show\nthat, at low bias and below Tc, the transfer of spin angular momentum pumped by an externally\ndriven ferromagnetic insulator is a ↵ected by the formation of Andreev bound states. These bound\nstates capture the essential physics of condensate-facilitated spin flow. For finite thicknesses of the\nsuperconducting layer, comparable to the coherence length, Andreev bound states are resonant and\ndominate the subgap spin current, rendering a significant enhancement in spin current over a wide\nrange of parameters. The resonant enhancement of the subgap transport channels establishes a\nprototype Fabry-P´ erot resonator for spin pumping.\nCombining an s-wave superconducting order, favoring\nelectrons to form a singlet state, with a ferromagnetic\norder, favoring spin alignment, leads to a powerful en-\nhancement or reduction of angular momentum transfer\nin spintronic devices [ 1]. This happens through various\nprocesses such as the formation of Andreev bound states\n(ABSs) [ 2–9], the generation of spin-triplet pairing [ 10–\n15], and crossed Andreev reflections (CARs) [ 16]. In this\nLetter, we present an experimentally feasible structure\nmanifesting spin current enhancement due to resonant\nABSs. We argue that the ABS resonances studied here\nmay be responsible for the magnetic damping increase\nin a superconducting hybrid multilayer reported in Ref.\n[17]\nTime-dependent ferromagnetic magnetization, when it\nis driven by an external rf field, generates spin pumping–\na spin angular momentum flow–into adjacent materials\nthat dissipates the energy of the ferromagnet [ 18]. Insert-\ning a thick superconductor (S) in between, however, can\nsuppress spin pumping due to the superconducting gap\nopening for quasiparticles (QPs). Consequently, in spin\ntransport studies with a bulk S, it is common to have\nbias voltages to inject QPs with energies above the gap\n[19]. Generically, for a thick superconducting layer, it is\nexpected that at low temperatures spin current relies on\na spin-polarized (triplet Cooper pairing) supercurrent in-\nduced at highly spin-active regions or complex magnetic\nmultilayers [ 20]. For a thin S layer (thickness ⇠coherence\nlength), on the other hand, subgap currents can be car-\nried via evanescent QPs tunneling through the S layer.\nIn this work, we study a hybrid structure consisting of\na thin S layer in proximity to a driven ferromagnetic in-\nsulator (FI) separated by a normal metal region (N). To\ncollect the spin current we have placed a spin reservoir\nNr, comprising an N r|S|N|FI structure, see Fig. 1. In or-\nder to establish the pumped spin current in the presence\nof superconducting and dynamic ferromagnetic orders,\nwe solve a time-dependent scattering problem [ 21,22],\nwhich accounts for the relevant processes such as AR,\nCAR, ABS, and dynamic triplet-paring generation. For\nsubgap energies, the full scattering matrix develops peakS!(#)\nFI%& xz\ny\nN' NdS \t\t dN \t\t\nFIG. 1. Schematic sketch for an N r|S|N|FI hybrid structure.\nThe FI region with precessing magnetization m(t) injects spin\nangular momentum that is carried by the QPs into the reser-\nvoir N r(spin pumping). To investigate the e ↵ect of supercon-\nductivity on the magnetically pumped spin flow, we utilize\nthe scattering formalism for incident and reflected QPs. For\nfinite thicknesses of the S and N layers ( dS,dN⇠coherence\nlength) the resonant ABSs provide highly transmitting trans-\nport channels for spin pumping. Here, for example, we show\nfour resonant channels by two positive-energy ABSs. The\nenergies are measured with respect to the Fermi energy EF\ndepicted in the middle.\nstructures marking resonant bound states (ABSs), which\nengender highly transmitting subgap transport channels\nfor the spin flow. Due to a resonant enhancement of spin\nflow for certain energies, as we will discuss, our setup can\nbe considered a prototype Fabry-P´ erot resonator for the\nspin pumping.\nThe spin pumping generated by the variations in the\nmagnetization direction m(t), assuming no voltage bias,\nis given by Is(t)=1\n4⇡g\"#\ne↵m⇥@tm[13,21,23], where the\ne↵ective spin-mixing conductance is defined as follows,\ng\"#\ne↵⌘Z1\n\u00001d\"@\"f(\")R e⇥\ng\"#⇤\n. (1)\nThe quantity Re⇥\ng\"#⇤\nrefers to the real part of the spin-\nmixing conductance g\"#=T r⇥\n1\u0000r\"\neer⇤#\nee+r#\"\nher⇤\"#\nhe⇤\n,w h e r e\nr\u0000\neeand r\u0000\u0000\u0000\nhe (\u00002{\",#}) represent the electron-to-The Supplemental Material\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nTHEORY OF A SPIN-DEPENDENT\nSCATTERING WITH SUPERCONDUCTOR\nIn the following, we study a time-dependent scatter-\ning in ballistic regime for a Nr|S|N|FIhybrid structure,\nsee Fig. 1. Our scattering process is carried out by spin-\nful QPs, i.e., electrons and holes, whose amplitudes are\ngiven by the complex numbers c±\ne,h.T h e s u p e r s c r i p t ±\nsign indicates the direction of the group velocity (normal\ncomponent) given by ±⇣e,hˆx, for which we have defined\n⇣e= +1, and ⇣h=\u00001. In the absence of magnetic in-\nhomogeneity and spin-orbit interaction, the spin is con-\nserved along magnetization m.\nIt is customary that at the interfaces, generic bound-\nary conditions pertinent to interface scattering and dis-\norder can be taken into account by an insertion of a per-\nfect spacer N between adjacent materials[ 1,2]. This way,\nthe full scattering process is broken down into studying\nNr|S|N, and N|FIstructures separately. We bear in\nmind that propagation through the N layer mixes nei-\nther particle-hole nor spin degrees of freedom. For the\nNr|S|Nstructure, following BTK[ 3], we assume a uni-\nform superconducting pair-potential \u0000inside the S layer\nthat vanishes in the N layers. This way the scattering\nprocess takes place only at the interfaces, which essen-\ntially contains QPs propagating into the S and Andreev\nretro-reflections. In a more realistic (dirty) normal metal-\nS interface, however, there is a possibility of a specular\nreflection for electrons and holes. Hence, to model par-\ntially reflective interfaces, we have incorporated barrier\npotentials in the form of Z1~vF\u0000(x), and Z2~vF\u0000(x\u0000dS)\nwhose strength are parametrized by Z1,Z2,(vFis the\nFermi velocity).\nInitially, we assume no barrier potentials and study\na single-channel scattering in a 1D wire. Physically, this\ncan be justified by imposing a constriction, that is, trans-\nverse thickness of the N layers to be on the order of a\nFermi wavelength[ 1]. Later on, we will relax these condi-\ntions by considering a 3D geometry with partially reflec-\ntive 2D interfaces ( Z1,Z26= 0), for which we integrate\nout all transverse modes.\nAs n a p s h o tc o n fi g u r a t i o ns c a t t e r i n gp r o c e s s\nA comparison between a typical FI precession period\n⇠10\u000010s (FMR period) with the time scales of elec-\ntron and hole dynamics, ( dS+dN)/vF⌧10\u000010s, re-\nveals that the magnetization dynamics can be taken as\nan adiabatic evolution during the scattering process. Ini-!\"#\n!\"$\n!%$\n!%#\nSm(t)\ndS \t\txz\n!\"#\n!\"$\n!%$\n!%#ke,h k∥\nFIN\nNr \t\tdN\t\t() (*\nFIG. 1. A snapshot configuration for magnetization m(t)i n\nanNr|S|N|FIstructure with incoming and outgoing modes\n(black and red, respectively). The propagation direction for\nelectrons and holes with scattering amplitudes c±\ne,c±\nhis given\nby±ˆxfor electrons, and ⌥ˆxfor holes. The shaded slabs at\nx=0a n d dSrepresent barrier potentials whose strengths are\nparametrized by Z1andZ2.T h e i n c i d e n t Q P s ’ w a v e v e c t o r\nis given by ( ke,h,kk).\ntially, this reduces the problem to a time-independent\nprocess carried out with a snapshot configuration for the\nmagnetization. One can generalize[ ?] to incorporate\nmagnetization precession dynamics with a simple SU(2)\nrotation to the rotating frame of m(t). Including the\nprecessional motion of the magnetization, which is rotat-\ning around z-axism(t)=\u0000\nsin✓cos!t,sin✓sin!t,cos✓\u0000\n,\ncan be accomplished by a spinor rotation[ ?] asS(\",t)=\nU†(t)S(\")U(t). The time-dependent unitary rotation\nU(t) is a transformation from the lab frame to rotating\nframe of m, that is given by\nU= \nei✓\n2\u0000y0\n0 ei✓\n2\u0000y!✓\nei!t\n2\u0000z 0\n0 e\u0000i!t\n2\u0000z◆\n. (1)\nThe precessional angular speed is given by !.\nTo start, let us begin with the scattering process in\nNr|S|Nstructure[ 4]. For a S with a thickness given by\ndS, applying continuity conditions for wave function and\ndiscontinuity of its derivative on the interfaces as well\nas current continuity inside the S, we can relate the in-\ncoming modes⇥\nc+\ne(Nr),c\u0000\ne(N),c\u0000\nh(Nr),c+\nh(N)⇤Tto the\noutgoing modes⇥\nc\u0000\ne(Nr),c+\ne(N),c+\nh(Nr),c\u0000\nh(N)⇤Tin the\nNr|S|Nstructure (for details of this calculation see Ap-\npendix ??).\nFor the normal metal Na width of dN, the electron\nand hole propagate through and reflect back o ↵the FI\ninterface without mixing particle-hole and spin spaces.The Supplemental Material\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nTHEORY OF A SPIN-DEPENDENT\nSCATTERING WITH SUPERCONDUCTOR\nIn the following, we study a time-dependent scatter-\ning in ballistic regime for a Nr|S|N|FIhybrid structure,\nsee Fig. 1. Our scattering process is carried out by spin-\nful QPs, i.e., electrons and holes, whose amplitudes are\ngiven by the complex numbers c±\ne,h.T h e s u p e r s c r i p t ±\nsign indicates the direction of the group velocity (normal\ncomponent) given by ±⇣e,hˆx, for which we have defined\n⇣e= +1, and ⇣h=\u00001. In the absence of magnetic in-\nhomogeneity and spin-orbit interaction, the spin is con-\nserved along magnetization m.\nIt is customary that at the interfaces, generic bound-\nary conditions pertinent to interface scattering and dis-\norder can be taken into account by an insertion of a per-\nfect spacer N between adjacent materials[ 1,2]. This way,\nthe full scattering process is broken down into studying\nNr|S|N, and N|FIstructures separately. We bear in\nmind that propagation through the N layer mixes nei-\nther particle-hole nor spin degrees of freedom. For the\nNr|S|Nstructure, following BTK[ 3], we assume a uni-\nform superconducting pair-potential \u0000inside the S layer\nthat vanishes in the N layers. This way the scattering\nprocess takes place only at the interfaces, which essen-\ntially contains QPs propagating into the S and Andreev\nretro-reflections. In a more realistic (dirty) normal metal-\nS interface, however, there is a possibility of a specular\nreflection for electrons and holes. Hence, to model par-\ntially reflective interfaces, we have incorporated barrier\npotentials in the form of Z1~vF\u0000(x), and Z2~vF\u0000(x\u0000dS)\nwhose strength are parametrized by Z1,Z2,(vFis the\nFermi velocity).\nInitially, we assume no barrier potentials and study\na single-channel scattering in a 1D wire. Physically, this\ncan be justified by imposing a constriction, that is, trans-\nverse thickness of the N layers to be on the order of a\nFermi wavelength[ 1]. Later on, we will relax these condi-\ntions by considering a 3D geometry with partially reflec-\ntive 2D interfaces ( Z1,Z26= 0), for which we integrate\nout all transverse modes.\nAs n a p s h o tc o n fi g u r a t i o ns c a t t e r i n gp r o c e s s\nA comparison between a typical FI precession period\n⇠10\u000010s (FMR period) with the time scales of elec-\ntron and hole dynamics, ( dS+dN)/vF⌧10\u000010s, re-\nveals that the magnetization dynamics can be taken as\nan adiabatic evolution during the scattering process. Ini-!\"#\n!\"$\n!%$\n!%#\nSm(t)\ndS \t\txz\n!\"#\n!\"$\n!%$\n!%#ke,h k∥\nFIN\nNr \t\tdN\t\t() (*\nFIG. 1. A snapshot configuration for magnetization m(t)i n\nanNr|S|N|FIstructure with incoming and outgoing modes\n(black and red, respectively). The propagation direction for\nelectrons and holes with scattering amplitudes c±\ne,c±\nhis given\nby±ˆxfor electrons, and ⌥ˆxfor holes. The shaded slabs at\nx=0a n d dSrepresent barrier potentials whose strengths are\nparametrized by Z1and Z2.T h e i n c i d e n t Q P s ’ w a v e v e c t o r\nis given by ( ke,h,kk).\ntially, this reduces the problem to a time-independent\nprocess carried out with a snapshot configuration for the\nmagnetization. One can generalize[ ?] to incorporate\nmagnetization precession dynamics with a simple SU(2)\nrotation to the rotating frame of m(t). Including the\nprecessional motion of the magnetization, which is rotat-\ning around z-axism(t)=\u0000sin✓cos!t,sin✓sin!t,cos✓\u0000,\ncan be accomplished by a spinor rotation[ ?] asS(\",t)=\nU†(t)S(\")U(t). The time-dependent unitary rotation\nU(t) is a transformation from the lab frame to rotating\nframe of m, that is given by\nU= \nei✓\n2\u0000y0\n0 ei✓\n2\u0000y!✓ei!t\n2\u0000z 0\n0 e\u0000i!t\n2\u0000z◆\n. (1)\nThe precessional angular speed is given by !.\nTo start, let us begin with the scattering process in\nNr|S|Nstructure[ 4]. For a S with a thickness given by\ndS, applying continuity conditions for wave function and\ndiscontinuity of its derivative on the interfaces as well\nas current continuity inside the S, we can relate the in-\ncoming modes⇥c+\ne(Nr),c\u0000\ne(N),c\u0000\nh(Nr),c+\nh(N)⇤Tto the\noutgoing modes⇥c\u0000\ne(Nr),c+\ne(N),c+\nh(Nr),c\u0000\nh(N)⇤Tin the\nNr|S|Nstructure (for details of this calculation see Ap-\npendix ??).\nFor the normal metal Na width of dN, the electron\nand hole propagate through and reflect back o ↵the FI\ninterface without mixing particle-hole and spin spaces.The Supplemental Material\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nTHEORY OF A SPIN-DEPENDENT\nSCATTERING WITH SUPERCONDUCTOR\nIn the following, we study a time-dependent scatter-\ning in ballistic regime for a Nr|S|N|FIhybrid structure,\nsee Fig. 1. Our scattering process is carried out by spin-\nful QPs, i.e., electrons and holes, whose amplitudes are\ngiven by the complex numbers c±\ne,h.T h e s u p e r s c r i p t ±\nsign indicates the direction of the group velocity (normal\ncomponent) given by ±⇣e,hˆx, for which we have defined\n⇣e= +1, and ⇣h=\u00001. In the absence of magnetic in-\nhomogeneity and spin-orbit interaction, the spin is con-\nserved along magnetization m.\nIt is customary that at the interfaces, generic bound-\nary conditions pertinent to interface scattering and dis-\norder can be taken into account by an insertion of a per-\nfect spacer N between adjacent materials[ 1,2]. This way,\nthe full scattering process is broken down into studying\nNr|S|N, and N|FIstructures separately. We bear in\nmind that propagation through the N layer mixes nei-\nther particle-hole nor spin degrees of freedom. For the\nNr|S|Nstructure, following BTK[ 3], we assume a uni-\nform superconducting pair-potential \u0000inside the S layer\nthat vanishes in the N layers. This way the scattering\nprocess takes place only at the interfaces, which essen-\ntially contains QPs propagating into the S and Andreev\nretro-reflections. In a more realistic (dirty) normal metal-\nS interface, however, there is a possibility of a specular\nreflection for electrons and holes. Hence, to model par-\ntially reflective interfaces, we have incorporated barrier\npotentials in the form of Z1~vF\u0000(x), and Z2~vF\u0000(x\u0000dS)\nwhose strength are parametrized by Z1,Z2,(vFis the\nFermi velocity).\nInitially, we assume no barrier potentials and study\na single-channel scattering in a 1D wire. Physically, this\ncan be justified by imposing a constriction, that is, trans-\nverse thickness of the N layers to be on the order of a\nFermi wavelength[ 1]. Later on, we will relax these condi-\ntions by considering a 3D geometry with partially reflec-\ntive 2D interfaces ( Z1,Z26= 0), for which we integrate\nout all transverse modes.\nAs n a p s h o tc o n fi g u r a t i o ns c a t t e r i n gp r o c e s s\nA comparison between a typical FI precession period\n⇠10\u000010s (FMR period) with the time scales of elec-\ntron and hole dynamics, ( dS+dN)/vF⌧10\u000010s, re-\nveals that the magnetization dynamics can be taken as\nan adiabatic evolution during the scattering process. Ini-!\"#\n!\"$\n!%$\n!%#\nSm(t)\ndS \t\txz\n!\"#\n!\"$\n!%$\n!%#ke,h k∥\nFIN\nNr \t\tdN\t\t() (*\nFIG. 1. A snapshot configuration for magnetization m(t)i n\nanNr|S|N|FIstructure with incoming and outgoing modes\n(black and red, respectively). The propagation direction for\nelectrons and holes with scattering amplitudes c±\ne,c±\nhis given\nby±ˆxfor electrons, and ⌥ˆxfor holes. The shaded slabs at\nx=0a n d dSrepresent barrier potentials whose strengths are\nparametrized by Z1andZ2.T h e i n c i d e n t Q P s ’ w a v e v e c t o r\nis given by ( ke,h,kk).\ntially, this reduces the problem to a time-independent\nprocess carried out with a snapshot configuration for the\nmagnetization. One can generalize[ ?] to incorporate\nmagnetization precession dynamics with a simple SU(2)\nrotation to the rotating frame of m(t). Including the\nprecessional motion of the magnetization, which is rotat-\ning around z-axism(t)=\u0000\nsin✓cos!t,sin✓sin!t,cos✓\u0000\n,\ncan be accomplished by a spinor rotation[ ?] asS(\",t)=\nU†(t)S(\")U(t). The time-dependent unitary rotation\nU(t) is a transformation from the lab frame to rotating\nframe of m, that is given by\nU= \nei✓\n2\u0000y0\n0ei✓\n2\u0000y!✓\nei!t\n2\u0000z 0\n0 e\u0000i!t\n2\u0000z◆\n. (1)\nThe precessional angular speed is given by !.\nTo start, let us begin with the scattering process in\nNr|S|Nstructure[ 4]. For a S with a thickness given by\ndS, applying continuity conditions for wave function and\ndiscontinuity of its derivative on the interfaces as well\nas current continuity inside the S, we can relate the in-\ncoming modes⇥\nc+\ne(Nr),c\u0000\ne(N),c\u0000\nh(Nr),c+\nh(N)⇤Tto the\noutgoing modes⇥\nc\u0000\ne(Nr),c+\ne(N),c+\nh(Nr),c\u0000\nh(N)⇤Tin the\nNr|S|Nstructure (for details of this calculation see Ap-\npendix ??).\nFor the normal metal Na width of dN, the electron\nand hole propagate through and reflect back o ↵the FI\ninterface without mixing particle-hole and spin spaces.The Supplemental Material\nMostafa Tanhayi Ahari, Yaroslav Tserkovnyak\nDepartment of Physics and Astronomy, University of California, Los Angeles, California 90095, USA\nTHEORY OF A SPIN-DEPENDENT\nSCATTERING WITH SUPERCONDUCTOR\nIn the following, we study a time-dependent scatter-\ning in ballistic regime for a Nr|S|N|FIhybrid structure,\nsee Fig. 1. Our scattering process is carried out by spin-\nful QPs, i.e., electrons and holes, whose amplitudes are\ngiven by the complex numbers c±\ne,h.T h e s u p e r s c r i p t ±\nsign indicates the direction of the group velocity (normal\ncomponent) given by ±⇣e,hˆx, for which we have defined\n⇣e= +1, and ⇣h=\u00001. In the absence of magnetic in-\nhomogeneity and spin-orbit interaction, the spin is con-\nserved along magnetization m.\nIt is customary that at the interfaces, generic bound-\nary conditions pertinent to interface scattering and dis-\norder can be taken into account by an insertion of a per-\nfect spacer N between adjacent materials[ 1,2]. This way,\nthe full scattering process is broken down into studying\nNr|S|N, and N|FIstructures separately. We bear in\nmind that propagation through the N layer mixes nei-\nther particle-hole nor spin degrees of freedom. For the\nNr|S|Nstructure, following BTK[ 3], we assume a uni-\nform superconducting pair-potential \u0000inside the S layer\nthat vanishes in the N layers. This way the scattering\nprocess takes place only at the interfaces, which essen-\ntially contains QPs propagating into the S and Andreev\nretro-reflections. In a more realistic (dirty) normal metal-\nS interface, however, there is a possibility of a specular\nreflection for electrons and holes. Hence, to model par-\ntially reflective interfaces, we have incorporated barrier\npotentials in the form of Z1~vF\u0000(x), and Z2~vF\u0000(x\u0000dS)\nwhose strength are parametrized by Z1,Z2,(vFis the\nFermi velocity).\nInitially, we assume no barrier potentials and study\na single-channel scattering in a 1D wire. Physically, this\ncan be justified by imposing a constriction, that is, trans-\nverse thickness of the N layers to be on the order of a\nFermi wavelength[ 1]. Later on, we will relax these condi-\ntions by considering a 3D geometry with partially reflec-\ntive 2D interfaces ( Z1,Z26= 0), for which we integrate\nout all transverse modes.\nAs n a p s h o tc o n fi g u r a t i o ns c a t t e r i n gp r o c e s s\nA comparison between a typical FI precession period\n⇠10\u000010s (FMR period) with the time scales of elec-\ntron and hole dynamics, ( dS+dN)/vF⌧10\u000010s, re-\nveals that the magnetization dynamics can be taken as\nan adiabatic evolution during the scattering process. Ini-!\"#\n!\"$\n!%$\n!%#\nSm(t)\ndS \t\txz\n!\"#\n!\"$\n!%$\n!%#ke,h k∥\nFIN\nNr \t\tdN\t\t() (*\nFIG. 1. A snapshot configuration for magnetization m(t)i n\nanNr|S|N|FIstructure with incoming and outgoing modes\n(black and red, respectively). The propagation direction for\nelectrons and holes with scattering amplitudes c±\ne,c±\nhis given\nby±ˆxfor electrons, and ⌥ˆxfor holes. The shaded slabs at\nx=0a n d dSrepresent barrier potentials whose strengths are\nparametrized by Z1andZ2.T h e i n c i d e n t Q P s ’ w a v e v e c t o r\nis given by ( ke,h,kk).\ntially, this reduces the problem to a time-independent\nprocess carried out with a snapshot configuration for the\nmagnetization. One can generalize[ ?] to incorporate\nmagnetization precession dynamics with a simple SU(2)\nrotation to the rotating frame of m(t). Including the\nprecessional motion of the magnetization, which is rotat-\ning around z-axism(t)=\u0000\nsin✓cos!t,sin✓sin!t,cos✓\u0000\n,\ncan be accomplished by a spinor rotation[ ?] asS(\",t)=\nU†(t)S(\")U(t). The time-dependent unitary rotation\nU(t) is a transformation from the lab frame to rotating\nframe of m, that is given by\nU= \nei✓\n2\u0000y0\n0 ei✓\n2\u0000y!✓\nei!t\n2\u0000z 0\n0 e\u0000i!t\n2\u0000z◆\n. (1)\nThe precessional angular speed is given by !.\nTo start, let us begin with the scattering process in\nNr|S|Nstructure[ 4]. For a S with a thickness given by\ndS, applying continuity conditions for wave function and\ndiscontinuity of its derivative on the interfaces as well\nas current continuity inside the S, we can relate the in-\ncoming modes⇥\nc+\ne(Nr),c\u0000\ne(N),c\u0000\nh(Nr),c+\nh(N)⇤Tto the\noutgoing modes⇥\nc\u0000\ne(Nr),c+\ne(N),c+\nh(Nr),c\u0000\nh(N)⇤Tin the\nNr|S|Nstructure (for details of this calculation see Ap-\npendix ??).\nFor the normal metal Na width of dN, the electron\nand hole propagate through and reflect back o ↵the FI\ninterface without mixing particle-hole and spin spaces.\nFIG. S1. A snapshot con\fguration for the magnetization m(t) in an N rjSjNjFI structure with the incoming and outgoing\nmodes into the S (black and red, respectively). The propagation direction for electrons and holes with scattering amplitudes\nc\u0006\ne;c\u0006\nhare given by\u0006^x,\u0007^x, respectively. The shaded slabs at x= 0 anddSrepresent the barrier potentials whose strengths\nare parametrized by Z1andZ2. The incident QP wave vector is given by ( ke;h;kk).\nTherefore, the incident and re\rected modes in N and S layers can be expanded as\n\tN=c+\ne(N)\t+\nNne+c\u0000\ne(N)\t\u0000\nNne+c+\nh(N)\t+\nNnh+c\u0000\nh(N)\t\u0000\nNnh; (S6)\nand\n\tS=c+\ne(S)\t+\nSne+c\u0000\ne(S)\t\u0000\nSne+c+\nh(S)\t+\nSnh+c\u0000\nh(S)\t\u0000\nSnh: (S7)\nThe scattering matrix across the S is obtained by imposing the following conditions: (1) continuity of the wave\nfunctions and their derivative at the interfaces x= 0;dS, (2) continuity of the current inside the S, that is, 0 0. This indicates that, in the limit of zero barrier, the insertion of a normal metal\nlayer not only provides a level of control but is a necessary element for the enhancement of spin \row. By increasing\nthe barrier potential strength, however, subgap normal bound states begin to form and ABSs are pushed away from\nthe subgap region (see Fig. S2). In order to see this, we have de\fned\nfABS(Z)\u0011Re g\"#(dS;dN;\"res;Z)\u0000Re g\"#(dS;dN;\"res;Z1): (S18)\nHere,\"resrepresent resonance energies, for which the real part of the conductance is maximum. Note that, \"res!\"A\nwhenZ!0. SinceZ1\u001d1, Re g\"#(dS;dN;\"res;Z1) includes only normal bound{state resonances.\n3D geometry .|Due to the translation symmetry on the interfaces, the full 3D scattering problem can be inferred\nfrom our scattering matrix given in Eq. (S16). As we have mentioned earlier, the wave number for the electron and\nhole propagating at energy \"is given by\nkn\ne;h=r\n2m\n~2p\nEF\u0000En+\u0010e;h\"; (S19)4\nFIG. S2. (a) The resonant bound{state energies, \"res, for various Z. WhenZ= 0, we get \"res=\"A. (b) For \fnite but\nsmall barrier strength, Re g\"#has contribution from normal and Andreev bound{state resonances. (c) For the stronger barrier,\nwhen the AR probability is negligible, the resonant energies are due to normal bound states. In (a), (b), and (c), we have set\ndS=\u00180= 1:5 and \u0001 ex= 0:4EF. (d)fABS(Z) is shown for Z1= 2, where the blue, orange, and green curves correspond to\ndS=\u00180= 1, 1:5, and 2, respectively. In this plot, we have \u0001 ex= 0:4EFanddN=\u00180= 1.\nFIG. S3. The normalized conductance averaged over N(\") = 20 channels in ballistic regime for various barrier strengths Z.\nHere, we have set \u0001 ex=EF= 0:1,T=Tc= 0:4, anddN=\u00180= 2:5.\nwhereEnis the transverse energy of the incident electron or hole. Here, the channel number index nis considered\nto distinguish QPs propagating through di\u000berent channels at energy \". The multichannel e\u000bective conductance is\nobtained by summing over all the channels present at energy \". We consider N(\") to denote the total number of\nsuch channels. In Fig. S3, we have shown normalized e\u000bective conductance for multichannel case in the presence of\ninterface barrier potential.\nDynamic generation of triplet Cooper pairs |We utilize a heuristic argument to describe the triplet pairing gen-\neration in our setup. For a thin S layer, dS\u0018\u00180, QPs with the resonant energies can escape the Nlayer either by5\ntunnelling through the S layer or forming Cooper pairs inside the S layer. For the latter case, the Cooper pairs contain\nthe spin{dependent phases as follows [6]:\nei\u000bj\"#i\u0000e\u0000i\u000bj#\"i; (S20)\nwhere\u000b\u0018#. These Cooper pairs contain a spin{triplet component, j\"#i+j#\"i, with an amplitude determined by sin \u000b.\nThe precessing magnetization mcan dynamically induce spin{polarized triplet Cooper pairs j##iandj\"\"i, which is\ncaused by the variations in the direction of the quantization axis for spins [6]. In other words, the spin{polarized\ntriplet components arise when the spin{triplet component j\"#i+j#\"iis written in a new rotated basis [6].\n[1] C. W. J. Beenakker. Rev. Mod. Phys. 69, 731 (1997)\n[2] Y. Nazarov, Y. Blanter. Quantum Transport: Introduction to Nanoscience. Cambridge: Cambridge University Press (2009)\n[3] G. E. Blonder, M. Tinkham, and T. M. Klapwijk. Phys. Rev. B 25, 4515 (1982)\n[4] O. Entin-Wohlman, Y. Imry, and A. Aharony. Phys. Rev. B 78, 224510 (2008)\n[5] H. J. Skadsem, A. Brataas, J. Martinek, and Y. Tserkovnyak, Phys. Rev. B 84, 104420 (2011)\n[6] M. Eschrig, T. L ofwander. Nature Physics, 4, (2008)\nJ. Linder, J. W. A. Robinson, Nature Physics, 11, (2015)" }, { "title": "2403.17976v1.Efecto_Hall_de_espin_inverso_en_peliculas_de_Nb_Mo_y_Bi_por_bombeo_de_espin.pdf", "content": " EFECTO HALL DE ESPÍN INVERSO EN PELÍCULAS DE Nb, Mo y Bi \nPOR BOMBEO DE ESPÍN \n \n \n \nD. Ley Domínguez1, J. A. Matutes -Aquino2 \n \n1Universidad Tecnológica de Ciudad Juárez, Av. Universidad Tecnológica # 3051.Col. Lote Bravo II, Cd. Juárez \nChihuahua, México , 32695 . \n2Centro de Investigación en Materiales Avanzados, S.C., Miguel de Cervantes 120, Complejo Industrial Chihuahua, \nChihuahua, México , 31109 . \nAutor Corresponsal : david_ley@utcj.edu.mx \n \n \nResumen: El efecto Hall de espín inverso utilizado para la detección de corrientes de espín fue observado mediante \nmedidas de voltaje en bicapas de metal normal (MN)/metal ferromagnético (MF), utilizando Nb, Mo y Bi como \nmetal normal y Permalloy (Py, Ni 81Fe19) como metal ferromagnético. La corriente de espín fue g enerada por el \nefecto de bombeo de espín con resonancia ferromagnética. Las muestras fueron depositadas por pulverización \ncatódica (sputtering) con magnetrón de corriente continua a temperatura ambiente sobre sustratos de Si (001). Las \ntres bicapas de Nb/P y, Mo/Py y Bi/Py tuvieron un acoplamiento espín -órbita lo suficientemente grande para poder \nobservar la generación de voltaje por efecto Hall de espín inverso . \n \nPalabras clave: Corriente de espín pura, efecto Hall de espín , efecto de bombeo de espín , resonancia \nferromagnética . \n \nAbstract : The inverse spin Hall effect used for detection of spin currents was observed by voltage measurements in \nbilayers of normal metal (NM)/ferromagne tic metal (FM), using Nb, Mo and Bi as normal metal and Permalloy (Py, \nNi81Fe19) as ferromagnetic metal. The spin current was generated by the spin pumping effect with ferromagnetic \nresonance. The samples were deposited by dc magnetron sputtering at room temperature on Si (001) substrates. The \nthree bilayers of Nb/Py, Mo/Py and Bi/Py had a spin -orbit coupling large enough to observe the voltage generation \nby spin Hall effect. \n \nKeywords: Pure spin current, spin Hall effect , spin pumping effect , ferromagnetic resonance . \n \n1. Introducción \nLa espintrónica es una disciplina que estudia una propiedad intrínseca de los electrones que es el \nespín, para mejorar la eficiencia y velocidad de los dispositivos electrónicos. Este nuevo tipo de \nelectrónica trabaja no solo con la carga de los electrones sino también c on su espín. Los \ndispositivos espintrónicos combinan la microelectrónica convencional con los efectos derivados \nde la interacción entre el espín del electrón y las propiedades magnéticas del material. El enfoque \npara utilizar el espín está basado en su ali neación (ya sea “arriba” o “abajo”) respecto a una \nreferencia (un campo magnético aplicado o la orientación de la magnetización d e una película \nferromagnética). Los dispositivos operan con cierta cantidad de corriente eléctrica que depende \nde una manera pr edecible de la dirección de alineación del espín. La implementación de la \nalineación de los espines a la electrónica convencional aña de a los dispositivos una mayor \ncapacidad y velocidad con bajo consumo de energía en los dispositivos electrónicos. Actualm ente \nlos discos duros de computadora y las memorias MRAM funcionan con dispositivos de \nespintrónica [ 1-3]. \n \nLos aspectos clave en la espintrónica son la generación y detección de corrientes de espín, en la \nelectrónica convencional la orientación del espín del electrón de la corriente eléctrica o corriente de carga está completamente al azar, como se esquematiza en la figura 1 a), en otras palabras, el \nespín del electrón no tiene ningún rol en los dispositivos electrónicos, por lo que se ignora por \ncompleto el espín del electrón. En contraste con la corriente eléctrica, la corriente de espines \npolarizados toma en consideración la orientación del espín pero también puede estar asociada con \nla corriente de carga, como se ilustra en la figura 1 b), o puede ser solo un flujo de espín sin flujo \nneto de carga, llamada corriente de espín pura , ilustrada en la figura 1 c). La corriente de espín \npura se puede entender como un flujo de electrones polarizados solo con espín hacia arriba (spin -\nup) sumánd ole a éste un flujo de electrones igual pero polarizados con espín hacia abajo (spin -\ndown) y fluyendo en dirección opuesta, por esta razón no hay un flujo neto de carga, la corriente \nde espín pura se puede ver como la diferencia entre el flujo de espín arr iba y espín abajo, en \ncontraste con la corriente de carga (corriente eléctrica) donde es la suma de los flujo de espín \narriba y espín abajo. \n \n \nFigura 1 .- Esquema del flujo de electrones: a) corriente eléctrica o corriente de carga, existe un flujo \nneto de carga y el espín de los electrones está orientado al azar, b) corriente de carga con espines \npolarizados, tiene un flujo neto de carga y la mayoría de los espines están orientados en la misma \ndirección, c) No tiene un flujo neto de carga \n0IIIc debido a que \n I I y tiene un flujo neto \nde espín \n0 2I IIIs . \n \nLas corrientes de espín pueden ser creadas por fenómeno s recientemente descubiertos como: ( i) \nEl efecto Hall de espín (SHE, siglas del inglés Spin Hall Effect ) como se esquematiza en la figura \n2, consiste en una acumulación de espín en los bordes del material en donde existe un flujo de \ncarga eléctrica donde los espines hacia arriba son desviados a un lado y los espines hacia abajo \nson desviados al otro lado del conductor, si la polaridad de la corr iente eléctrica cambia, la \norientación de los espines también cambia. \n \n \nFigura 2. - Esquema del efecto Hall de espín (SHE) y efecto Hall de espín inverso (ISHE), a) una \ncorriente de carga con espines no polarizados fluye a través de un material no magnéti co con \nacoplamiento espín -órbita generando una corriente de espín pura transversal a ésta, b) una corriente de \nespín pura fluyendo en un material no magnético con acoplamiento espín -órbita genera una corriente \neléctrica transversal. \n \nEl efecto Hall de esp ín en metales no magnéticos (MN) se origina en el acoplamiento espín -\nórbita, acoplamiento entre el espín del electrón y su movimiento orbital, generalmente encontrado \nen átomos grandes con un número atómico grande. De manera contraria , si una corriente de espín \npura fluye en un material no magnético que tenga fuerte acoplamiento espín -órbita generará un \nflujo de corriente de carga transversal a la corriente de espín, este efecto es llamado efecto Hall \nde espín inverso (ISHE, siglas del inglés Inverse Spin H all Effect ) [4-8]. (ii) El efecto de bombeo \nde espín (SPE, siglas del inglés Spin Pumping Effect), en donde una corriente de espín pura \npuede ser inyectada en un material no magnético debido a la precesión de un ma terial \nferromagnético adyacente. L a precesión de la dirección de la magnetización en el material \nferromagnético debido al torque ejercido por un campo de RF (radio frecuencia) externo bombea \nespines dentro del metal normar adyacente induciendo una corriente de espín pura [9-12]. Los \nefectos SHE e ISHE son requeridos para la creación de dispositivos espintrónicos siendo el \nPlatino uno de los elementos más utilizado en el estudio de estos efectos [13-14]. Debido al alto \ncosto del platino es importante buscar otras alternativas. En este artículo se reportan medidas de \nvoltaje obtenido del efecto Hall de espín inverso por la corriente de espín generada mediante el \nefecto de bombeo de espín para los metales de Nb, Mo y Bi. \n \n2. Experimentación \nLas películas fueron depositadas por la técnica de pulverización catódica (sputtering) con \nmagnetrón de corriente continua a temperatura ambiente, se utilizaron substratos de silicio \nmonocristalino con orientación (001) con un espesor de 500 µm previamente lavados con \nultrasonido , se cortaron en rectángulos de 1 mm x 3 mm . Se utilizaran mascaras para depositar \nlos metales normales en el centro de la muestra con una dimensión de 1 mm x 1 mm . La cámara \nde pulverización fue evacuada a una presión base de 2.2 x 10-7 Torr, después fue presurizada con \nun flujo de argón a 2.4 x 10-3 Torr durante el depósito. Se utilizaron blancos de 2 pulgadas de \ndiámetro de Permaloy (Ni 81Fe19), Niobio , Molibdeno y Bismuto, éstos con una pureza de \n99.95% . Todas las bicapas M N/MF de las muestras fueron fabricadas con un espesor de Py de 12 \nnm y un tiempo de depósito de 1 minuto a 50 mA para las películas de MN. Las muestras fueron \npuestas en resonancia ferromagnética con un espectrómetro de banda -X operado a una frecuencia \nde 9.4G Hz, la muestra fue posicionada en el centro de una cavidad resonante. Las muestras \nfueron barridas con un campo magnético externo generado por un electroimán de 9 pulgadas, \naplicado en dirección del plano de las películas . El voltaje generado como resultad o del efecto \nHall de espín inverso se midió con un nanovoltímetro conectado mediante una interface a una \ncomputadora. En la figura 3 se observa un esquema de la muestra. \n \n \nFigura 3. - Representación de la muestra de interface metal -normal /permalloy . \n \n3. Resultados y discusión \nEn la figura 4 (a -c) se observa el voltaje detectado por efecto Hall de espín inverso para l as \nmuestras de Nb/Py, Mo/Py y Bi/Py . En las tres bicapas se observa un pico de voltaje en \naproximadamente 1.1kOe, que es el campo magnético de resonancia del Py para una frecuencia \nde 9.4GHz. \nEl pico de voltaje observado para las tres muestras se encuentra en el rango de los µV , valores \nsimil ares han sido reportados para interfaces de Py/Py [ 13-14]. El voltaje se genera cuando la s \nbicapa s Nb/Py, Mo/Py y Bi/Py están en resonancia ferromagnética, una corriente de espín pura \nes inyectada de la película fer romagnética a la película de MN por medio del efecto de bombeo \nde espín. Una vez que existe una corriente de espín pura dentro de la película de MN se genera \nuna acumulación de cargas en las extremidades de la muestra por el desvío de electrones debido \nal acoplamiento espín -órbita , obteniendo un voltaje de efecto Hall de espín inverso . \nLa forma del pico de voltaje Hall de espín inverso es la misma que el pico de absorción de \nresonancia ferromagnética debido a que se crea cuando la película ferromagnética Py se \nencuentra en un campo magnético y fre cuencia adecuada cumpliendo con las condiciones de \nresonancia, estando el Py en resonancia ferromagnética la inyección de corriente de espín hacia \nlas películas de Nb, Mo y Bi es máxima, obteniendo el máximo voltaje. Una vez saliendo de las \ncondiciones de resonancia el bombeo de espín desaparece, debido a que fuera de las condiciones \nde resonancia magnética no existe una corriente de espín pura bombeada a la película de MN , el \nefecto Hall de espín inverso no se acciona y no genera una diferencia de potencial. \n \n \nFigura 4 .- Voltaje detectado por efecto Hall de espín inverso (ISHE) generado por la corriente de espín \ncreada mediante el bombeo de espín de resonancia ferromagnética, para las bicapas (a) Nb/Py, (b) \nMo/Py y (c) Bi/Py . \n \n4. conclusión \n Se obtuvieron medidas de voltaje por efecto Hall de espín inverso en bicapas de Nb/Py, Mo/Py y \nBi/Py . Por medio de las medidas de voltaje observadas en los resultados se pudo comprobar que \nlos metales Nb, Mo y Bi tienen un acoplamiento espín -órbita lo sufi cientemente grande para \nobservar el efecto Hall de espín inverso, por lo que pueden ser utilizados como generadores y \ndetectores de corrientes de espín . Debido al alto costo del Platino, s aber que el Nb, Mo y Bi \ngeneran voltaje por efecto Hall de espín inverso mediante bombeo de espín es una gran \nalternativa para minimizar costos en el área de la espintrónica . Estos elementos pueden ser \nestudiados para desarrollar nuevos dispositivos espintrónicos. \n \nReferencias \n1. Pulizzi F, Sinova J, Žutić I. (2012). Spintronics. Nature materials insight , 11, 367 -371. ISSN #1476 -1122. \n2. Žutić I, Jaroslav Fabian, Das Sarma S. (2004). Spintronics: Fundamentals and applications. Rev. Mod. Phys. 76, \n323.ISSN #1539 -0756. \n3. Wolf S.A , Awschalom D.D, Buhrman R.A, Daughton J.M, Molnár S. Von , Roukes M.L, Chtchelkanova A.Y, \nTreger D.M. (2001). Spintronics: A Spin -Based Electronics Vision for the Future. Science , 294, 1488 -1495. ISSN \n#1095 -9203. \n4. Dyakonov M. I, Perel V.I. (1971). Current -induced spin orientation of electrons in semiconductors. Physics \nletters A, 35, 459 -460. ISSN #0375 -9601. \n5. Hirsch J.E. (1999). Spin Hall Effect. Physical Review Letters , 83, 1834 . ISSN #1079 -7114. \n6. Kato Y.K, Myers R.C, Gossard A.C, Awschalom D.D. (2004). Observation of the spin Hall effect in \nsemiconductors. Science , 306, 1910. ISSN #1095 -9203. \n7. Takahashi Saburo, Maekawa Sadamichi. (2008), Spin current, spin accumulation and spin Hall effect. Sci. \nTechnol. Adv. Mater . 9, 014105. ISSN #1878 -5514. \n8. Vignale G. (2009). Ten Years of Spin Hall Effect. Journal of Superconductivity and Novel Magnetism , 23, ISSN \n#1557 -1947. \n9. Tserkovnyak Yaroslav, Brataas Arne, Bauer G. E. W. (2002). Spin pumping and magnetization dynamics in \nmetallic multilayers. Physi cal Review B , 66, 224403. ISSN #2469 -9969. \n10. Tserkovnyak Yaroslav, Brataas Arne, Bauer G. E. W. (2002). Enhanced Gilbert Damping in Thin Ferromagnetic \nFilms. Physical review Letters , 88, 117601. ISSN #1079 -7114. \n11. Azevedo A, Vilela -Leao L. H , Rodríguez -Suárez R. L , Lacerda Santos A. F , Rezende S. M. (2011). Spin \npumping and anisotropic magnetoresistance voltages in magnetic bilayers: Theory and experiment. Physical \nReview B , 83, 144402. ISSN #2469 -9969. \n12. Azevedo A, Vilela -Leao L. H, Rodríguez -Suárez R. L, Oliveira A. B, Rezende S. M. (2005). dc effect in \nferromagnetic resonance: Evidence of the spin -pumping effect?. Journal of Applied Physics , 97, 10C715. ISSN \n#1089 -7550. \n13. Nakayama H, Ando K, Harii K, Kajiwara Y, Yoshino T, Uchida K, Ota T, Saitoh E. (2010). Detection or inverse \nspin-Hall effect induced in Pt x-1Mx(M = Cu, Au) thin films. Journal of Physics: Conference Series , 200, 062014. \nISSN #1742 -6596. \n14. Nakayama H, Ando K, Harii K, Fujikawa Y, Kaj iwara Y, Yoshino T, Saitoh E. (2011). Inverse spin -Hall effect \ninduced by spin pumping in different size Ni 81Fe19/Pt films. Journal of Physics: Conference Series , 266, 012100. \nISSN #1742 -6596. \n " }, { "title": "1803.10925v1.Giant_resonant_nonlinear_damping_in_nanoscale_ferromagnets.pdf", "content": "Giant resonant nonlinear damping in nanoscale ferromagnets\nI. Barsukov,1,\u0003H. K. Lee,1A. A. Jara,1Y.-J. Chen,1A. M. Gon\u0018 calves,1\nC. Sha,1J. A. Katine,2R. E. Arias,3B. A. Ivanov,4, 5and I. N. Krivorotov1\n1Physics and Astronomy, University of California, Irvine, CA 92697, USA\n2Western Digital, 5600 Great Oaks Parkway, San Jose, CA 95119, USA\n3Departamento de F\u0013 \u0010sica, CEDENNA, FCFM, Universidad de Chile, Santiago, Chile\n4Institute of Magnetism, National Academy of Sciences of Ukraine, Vernadsky av. 36 B, Kyiv, 03142, Ukraine\n5National University of Science and Technology MISiS, Moscow, 119049, Russian Federation\nMagnetic damping is a key metric for emerging technologies based on magnetic nanoparticles,\nsuch as spin torque memory and high-resolution biomagnetic imaging. Despite its importance,\nunderstanding of magnetic dissipation in nanoscale ferromagnets remains elusive, and the damping\nis often treated as a phenomenological constant. Here we report the discovery of a giant frequency-\ndependent nonlinear damping that strongly alters the response of a nanoscale ferromagnet to spin\ntorque and microwave magnetic \feld. This novel damping mechanism originates from three-magnon\nscattering that is strongly enhanced by geometric con\fnement of magnons in the nanomagnet. We\nshow that the giant nonlinear damping can invert the e\u000bect of spin torque on a nanomagnet leading\nto a surprising current-induced enhancement of damping by an antidamping torque. Our work\nadvances understanding of magnetic dynamics in nanoscale ferromagnets and spin torque devices.\nI. INTRODUCTION\nNanoscale magnetic particles are the core components\nof several emerging technologies such as nonvolatile spin\ntorque memory [1], spin torque oscillators [2{7], targeted\ndrug delivery, and high-resolution biomagnetic imaging\n[8{11]. Control of magnetic damping holds the key to\nimproving the performance of many nanomagnet-based\npractical applications. In biomagnetic characterization\ntechniques such as magnetic resonance imaging [12], re-\nlaxometry [13], and magnetic particle imaging [14, 15],\nmagnetic damping a\u000bects nanoparticles relaxation times\nand image resolution. In spin torque memory and oscil-\nlators, magnetic damping determines the electrical cur-\nrent necessary for magnetic switching [1] and generation\nof auto-oscillations [16] and thereby determines energy-\ne\u000eciency of these technologies. The performance of\nnanomagnet-based microwave detectors is also directly\na\u000bected by the damping [17{19]. Despite its impor-\ntance across multiple disciplines, magnetic damping in\nnanoparticles is poorly understood and is usually mod-\neled as a phenomenological constant [6, 16].\nIn this article, we experimentally demonstrate that a\nferromagnetic nanoparticle can exhibit dynamics quali-\ntatively di\u000berent from those predicted by the constant\ndamping model. We show that nonlinear contributions\nto the damping can be unusually strong and the damp-\ning parameter itself can exhibit resonant frequency de-\npendence. Our work demonstrates that nonlinear damp-\ning in nanomagnets is qualitatively di\u000berent from that in\nbulk ferromagnets and requires a new theoretical frame-\nwork for its description. We show both experimentally\nand theoretically that such resonant nonlinear damping\noriginates from multi-magnon scattering in a magnetic\n\u0003igorb@ucr.edusystem with a discrete spectrum of magnons induced by\ngeometric con\fnement.\nWe also discover that the resonant nonlinear damping\ndramatically alters the response of a nanomagnet to spin\ntorque. Spin torque arising from injection of spin cur-\nrents polarized opposite to the direction of magnetization\nacts as negative damping [2]. We \fnd, however, that the\ne\u000bect of such antidamping spin torque is reversed, lead-\ning to an enhanced dissipation due to the nonlinear res-\nonant scattering. This counterintuitive behavior should\nhave signi\fcant impact on the operation of spin torque\nbased memory [1], oscillators [2{7] and microwave detec-\ntors [17{19].\nII. RESULTS\nA. Spin wave spectroscopy\nWe study nonlinear spin wave dynamics in nanoscale\nelliptical magnetic tunnel junctions (MTJs) that consist\nof a CoFeB free layer (FL), an MgO tunnel barrier, and a\nsynthetic antiferromagnet (SAF) pinned layer [20]. Spec-\ntral properties of the FL spin wave modes are studied in a\nvariety of MTJs with both in-plane and perpendicular-to-\nplane equilibrium orientations of the FL and SAF magne-\ntization. We observe strong resonant nonlinear damping\nin both the in-plane and the perpendicular MTJs, which\npoints to the universality of the e\u000bect.\nWe employ spin torque ferromagnetic resonance (ST-\nFMR) to measure magnetic damping of the FL spin wave\nmodes. In this technique, a microwave drive current\nIacsin(2\u0019ft) applied to the MTJ excites oscillations of\nmagnetization at the drive frequency f. The resulting\nmagnetoresistance oscillations Racsin(2\u0019ft+\u001e) generate\na direct voltage Vmix. Peaks in ST-FMR spectra Vmix(f)\narise from resonant excitation of spin wave eigenmodes\nof the MTJ [21{28]. To improve signal-to-noise ratio,arXiv:1803.10925v1 [cond-mat.mes-hall] 29 Mar 20182\n0 0.5 1 1.5 2H1H2\n36912\nField (kOe)0 0.5 1 1.5 2H2H1\n00.20.4\nField (kOe)Linewidth (GHz)Experiment\nSimulationFrequency (GHz)\n+1\n1a b\n~\nFIG. 1. Spin wave spectra in a nanoscale MTJ. (a) Normalized ST-FMR spectra h~Vmix(f)iof spin wave eigenmodes in a\nperpendicular MTJ device (Sample 1) measured as a function of out-of-plane magnetic \feld. Resonance peaks arising from\nthree low frequency modes of the MTJ free layer j0i,j1i, and j2iare observed. (b) Spectral linewidth of the quasi-uniform\nj0ispin wave mode as a function of out-of-plane magnetic \feld. Strong linewidth enhancement is observed in the resonant\nthree-magnon regime at H1andH2.\nthe magnitude of external magnetic \feld Happlied par-\nallel to the free layer magnetization is modulated, and\na \feld-derivative signal ~Vmix(f) = dVmix(f)=dHis mea-\nsured via lock-in detection technique [20]. Vmix(f) can\nthen be obtained via numerical integration (Supplemen-\ntal Material).\nFigure 1(a) shows ST-FMR spectra ~Vmix(f) measured\nas a function of out-of-plane magnetic \feld Hfor an el-\nliptical 52 nm\u000262 nm perpendicular MTJ device (Sam-\nple 1). Three spin wave eigenmodes with nearly linear\nfrequency-\feld relation fn(H) are clearly visible in the\nspectra. Micromagnetic simulations (Supplemental Ma-\nterial) reveal that these modes are three lowest frequency\nspin wave eigenmodes of the FL (Supplemental Material).\nThe lowest frequency (quasi-uniform) mode j0iis node-\nless and has spatially uniform phase. Each of the two\nhigher-order modes jni(n= 1;2) has a single node at\nthe FL center that is either perpendicular ( n= 1) or\nparallel (n= 2) to the ellipse long axis.\nThe spectral linewidth of the resonances in Fig. 1(a)\ncan be used for evaluation of the mode damping. The\nquasi-uniform mode j0iresonance visibly broadens at\ntwo magnetic \feld values: H1= 0:74 kOe (4 GHz) and\nH2= 1:34 kOe (6 GHz). Near H1, the modej1iresonance\nalso broadens and exhibits splitting, same behavior is ob-\nserved for the mode j2iatH2. At these \felds, the higher-\norder mode frequency is twice that of the quasi-uniform\nmodefn= 2f0. This shows that three-magnon con\ru-\nence [29{33] is the mechanism of the quasi-uniform mode\ndamping increase: two magnons of the quasi-uniform\nmodej0imerge into a single magnon of the higher-ordermodejni.\nThe most striking feature of the quasi-uniform mode\nresonance near H1is its split-peak shape with a local min-\nimum at the resonance frequency. Such a lineshape can-\nnot be \ft by the standard Lorentzian curve with symmet-\nric and antisymmetric components [20]. We therefore use\na double-peak \ftting function (Supplemental Material)\nto quantify the e\u000bective linewidth \u0001 f0of the resonance\npro\fle. For applied \felds su\u000eciently far from H1, the\nST-FMR curve recovers its single-peak shape and \u0001 f0\nis determined as half width of the standard Lorentzian\n\ftting function [20]. Figure 1(b) shows \u0001 f0as a function\nofHand demonstrates a large increase of the linewidth\nnear the \felds of the resonant three-magnon regime H1\nandH2. The stepwise increase of \u0001 f0nearH1is a result\nof the ST-FMR curve transition between the split-peak\nand single-peak shapes. For \felds near H2, the resonance\npro\fle broadens but does not develop a visible split-peak\nlineshape. As a result, \u0001 f0(H) is a smooth function in\nthe vicinity of H2.\nB. E\u000bect of spin torque\nIn MTJs, direct bias current Idcapplied across the\njunction exerts spin torque on the FL magnetization, act-\ning as antidamping for Idc>0 and as positive damping\nforIdc<0 [22, 34]. The antidamping spin torque in-\ncreases the amplitude of the FL spin wave modes [22, 35]\nand decreases their spectral linewidth [36]. We can em-\nploy spin torque from Idcto control the amplitude of spin3\nFIG. 2. E\u000bect of spin torque on spin wave resonance lineshape. (a)-(b) Spin wave resonance lineshapes in the nonresonant\nregime at H > H 1for di\u000berent values of direct bias current Idc. (c)-(d) Spin wave resonance lineshapes in the resonant three-\nmagnon regime at H=H1. (a), (c) Measured ST-FMR spectra (Sample 2). (b), (d) Solutions of Eqs. (3) and (4). Identical\nbias current values Idc(displayed in (a) are used in (a)-(d).\nwave eigenmodes excited in ST-FMR measurements, and\nthereby study the crossover between linear and nonlinear\nregimes of spin wave resonance.\nFigure 2 shows the dependence of ST-FMR resonance\ncurve of thej0imodeVmix(f) onIdcfor a 50 nm\u0002110 nm\nelliptical in-plane MTJ (Sample 2). For in-plane mag-\nnetic \feld values far from the three-magnon resonance\n\feldsHn, the amplitude of ST-FMR resonance curve\nVmix(f) shown in Fig. 2(a) monotonically increases with\nincreasing antidamping spin torque, as expected. At\nH=H1, the antidamping spin torque has a radically\ndi\u000berent and rather surprising e\u000bect on the resonance\ncurve. As illustrated in Fig. 2(c), increasing antidamp-\ning spin torque \frst broadens the resonance at H=H1\nand then transforms a single-peak resonance lineshape\ninto a split-peak lineshape with a local minimum at the\nresonance frequency f0. The data in Fig. 2 demonstrate\nthat the unusual split-peak lineshape of the resonance is\nonly observed when (i) the three-magnon scattering of\nthe quasi-uniform mode is allowed by the conservation of\nenergy and (ii) the amplitude of the mode is su\u000eciently\nhigh, con\frming that the observed e\u000bect is resonant and\nnonlinear in nature.\nFig. 2(c) reveals that antidamping spin torque can in-\ncrease the spectral linewidth and the e\u000bective damping\nof the quasi-uniform spin mode if the mode undergoes\nresonant three-magnon scattering. Figure 3 further illus-\ntrates this counterintuitive e\u000bect. It shows the linewidth\nof the quasi-uniform mode of a 50 nm \u0002110 nm elliptical\nin-plane MTJ (Sample 3) measured as a function of bias\ncurrent. In Fig. 3, blue symbols show the linewidth mea-\nsured at an in-plane magnetic \feld su\u000eciently far fromthe three-magnon resonance \felds Hn. At this \feld, the\nexpected quasi-linear dependence of the linewidth on Idc\nis observed for currents well below the critical current\nfor the excitation of auto-oscillatory magnetic dynamics.\nNear the critical current, the linewidth increases due to\na combination of the fold-over e\u000bect [37{39] and ther-\nmally activated switching between the large- and small-\namplitude oscillatory states of the fold-over regime [22].\nThe red symbols in Fig. 3 show the linewidth measured\nin the resonant three-magnon regime at H=H1. In con-\ntrast to the nonresonant regime, the linewidth increases\nwith increasingjIdcjfor both current polarities. Fur-\nthermore, the maximum linewidth is measured for the\nantidamping current polarity.\nIII. THEORETICAL MODEL\nNonlinear interactions among spin wave eigenmodes\nof a ferromagnet give rise to a number of spectacu-\nlar magneto-dynamic phenomena such as Suhl instabil-\nity of the uniform precession of magnetization [40, 41],\nspin wave self-focusing [42] and magnetic soliton forma-\ntion [43{45]. In bulk ferromagnets, nonlinear interac-\ntions generally couple each spin wave eigenmode to a\ncontinuum of other modes via energy- and momentum-\nconserving multi-magnon scattering [40]. This kinemat-\nically allowed scattering limits the achievable amplitude\nof spin wave modes and leads to broadening of the spin\nwave resonance. These processes lead to a resonance\nbroadening [40, 46{48] and cannot explain the observed\nsplit-peak lineshape of the resonance. In nanoscale ferro-4\nmagnets, geometric con\fnement discretizes the spin wave\nspectrum and thereby generally eliminates the kinemati-\ncally allowed multi-magnon scattering. This suppression\nof nonlinear scattering enables persistent excitation of\nspin waves with very large amplitudes [49] as observed in\nnanomagnet-based spin torque oscillators [2, 50]. Tun-\nability of the spin wave spectrum by external magnetic\n\feld, however, can lead to a resonant restoration of the\nenergy-conserving scattering [31]. The description of\nnonlinear spin wave resonance in the nanoscale ferromag-\nnet geometry therefore requires a new theoretical frame-\nwork. To derive the theory of resonant nonlinear damp-\ning in a nanomagnet, we start with a model Hamilto-\nnian that explicitly takes into account resonant nonlinear\nscattering between the quasi-uniform mode and a higher-\norder spin wave mode (in reduced units with ~\u00111):\nH=!0aya+!nbyb+\t0\n2ayayaa+\tn\n2bybybb (1)\n+( naaby+ \u0003\nnayayb)\n+\u0010\b\nexp(\u0000i!t)ay+ exp(i!t)a\t\nwhereay,aandby,bare the magnon creation and an-\nnihilation operators for the quasi-uniform mode j0iwith\nfrequency!0and for the higher-order spin wave mode\njnimode with frequency !n, respectively. The non-\nlinear mode coupling term proportional to the coupling\nstrength parameter ndescribes the annihilation of two\nj0imagnons and creation of one jnimagnon, as well as\nthe inverse process. The Hamiltonian is written in the\nresonant approximation, where small nonresonant terms\nsuch asaab,aaayare neglected. The terms proportional\nto \t 0and \t ndescribe the intrinsic nonlinear frequency\nshifts [51] of the modes j0iandjni. The last term de-\nscribes the excitation of the quasi-uniform mode by an\nexternal ac drive with the amplitude \u0010and frequency !.\nWe further de\fne classically a dissipation function Q,\nwhere\u000b0and\u000bnare the intrinsic linear damping param-\neters of the modes j0iandjni[52{54]:\nQ=day\ndtda\ndt(\u000b0+\u00110aya) +dby\ndtdb\ndt(\u000bn+\u0011nbyb) (2)\nFor generality, Eq. (2) includes intrinsic nonlinear\ndamping [16] of the modes j0iandjnidescribed by the\nnonlinearity parameters \u00110and\u0011n. However, our analy-\nsis below shows that the split-peak resonance lineshape\nis predicted by our theory even if \u00110and\u0011nare set equal\nto zero.\nEquations describing the nonlinear dynamics of the\ntwo coupled spin wave modes of the system follow from\nEq. (1) and Eq. (2):\nida\ndt=@H\n@ay+@Q\n@(day=dt)(3)\nidb\ndt=@H\n@by+@Q\n@(dby=dt)(4)\nIt can be shown (Supplemental Material) that these\nequations have a periodic solution a= \u0016aexp (\u0000i!t) and\n100 50 0 50 10000.10.20.30.40.5\nI (A)f0 (GHz)Resonant regime\nNonresonant regimeFIG. 3. E\u000bect of spin torque on linewidth. Linewidth of the\nquasi-uniform spin wave mode as a function of the applied\ndirect bias current (Sample 3): blue symbols { in the non-\nresonant regime H6=H1and red symbols { in the resonant\nthree-magnon regime H=H1. Lines are numerical \fts using\nEqs. (3) and (4).\nb=\u0016bexp (\u0000i2!t), where \u0016a,\u0016bare the complex spin wave\nmode amplitudes. For such periodic solution, Eqs. (3)\nand (4) are reduced to a set of two nonlinear algebraic\nequations for absolute values of the spin wave mode am-\nplitudesj\u0016ajandj\u0016bj, which can be solved numerically.\nSince the ST-FMR signal is proportional to j\u0016aj2(Supple-\nmental Material), the calculated j\u0016aj2(!) function can be\ndirectly compared to the measured ST-FMR resonance\nlineshape.\nWe employ the solution of Eqs. (3) and (4) to \ft the\n\feld dependence of the quasi-uniform mode linewidth in\nFig. 1(b). In this \ftting procedure, the resonance line-\nshapej\u0016aj2(!) is calculated, and its spectral linewidth\n\u0001!0is found numerically. The resonance frequencies !0\nand!nare directly determined from the ST-FMR data\nin Fig. 1(a). The intrinsic damping parameters \u000b0and\n\u000bnnearH1andH2are found from linear interpolations\nof the ST-FMR linewidths \u0001 f0and \u0001fnmeasured at\n\felds far from H1andH2. We \fnd that \u0001 !0weakly\ndepends on the nonlinearity parameters \t and \u0011, and\nthus these parameters are set to zero (Supplemental Ma-\nterial). We also \fnd that the calculated linewidth \u0001 !0\ndepends on the product of the drive amplitude \u0010and\nmode coupling strength n, but is nearly insensitive to\nthe individual values of \u0010and nas long as\u0010\u0001 n= const\n(Supplemental Material). Therefore, we use \u0010\u0001 nas a\nsingle \ftting parameter in this \ftting procedure. Solid\nline in Fig. 1(b) shows the calculated \feld dependence\nof the quasi-uniform mode linewidth on magnetic \feld.\nThe agreement of this single-parameter \ft with the ex-\nperiment is excellent.\nFigures 2(b) and 2(d) illustrate that Eqs. (3) and\n(4) not only describe the \feld dependence of ST-FMR\nlinewidth but also qualitatively reproduce the spectral5\nlineshapes of the measured ST-FMR resonances as well\nas the e\u000bect of the antidamping spin torque on the line-\nshapes. Fig. 2(b) shows the dependence of the calculated\nlineshapej\u0016aj2(!) on antidamping spin torque for a mag-\nnetic \feldHfar from the three-magnon resonance \felds\nHn. At this nonresonant \feld, increasing antidamping\nspin torque induces the fold-over of the resonance curve\n[37] without resonance peak splitting. The dependence of\nj\u0016aj2(!) on antidamping spin torque for H=H1is shown\nin Fig. 2(d). At this \feld, the resonance peak in j\u0016aj2(!)\n\frst broadens with increasing antidamping spin torque\nand then splits, in qualitative agreement with the ex-\nperimental ST-FMR data in Fig. 2(c). Our calculations\n(Supplemental Material) reveal that while the nonlinear-\nity parameters \t 0,\u00110, \tnand\u0011nhave little e\u000bect on\nthe linewidth \u0001 !0, they modify the lineshape of the res-\nonance. Given that the nonlinearity parameter values\nare not well known for the systems studied here, we do\nnot attempt to quantitatively \ft the measured ST-FMR\nlineshapes.\nEquations (3) and (4) also quantitatively explain\nthe observed dependence of the quasi-uniform mode\nlinewidth \u0001 !0on direct bias current Idc. Assuming an-\ntidamping spin torque linear in bias current [36, 55, 56]:\n\u000b0!\u000b0(1\u0000Idc=Ij0i\nc),\u000bn!\u000bn(1\u0000Idc=Ijni\nc), where\nIjni\nc>Ij0i\ncare the critical currents, we \ft the measured\nbias dependence of ST-FMR linewidth in Fig. 3 by solv-\ning Eqs. (3) and (4). The solid lines in Fig. 3 are the best\nnumerical \fts, where \u0010\u0001 nandIcare used as indepen-\ndent \ftting parameters. The rest of the parameters in\nEqs. (3) and (4) are directly determined from the experi-\nment following the procedure used for \ftting the data in\nFig. 1(b). Theoretical curves in Fig. 3 capture the main\nfeature of the data at the three-magnon resonance \feld\nH1{ increase of the linewidth with increasing antidamp-\ning spin torque.\nIV. DISCUSSION\nFurther insight into the mechanisms of the nonlinear\nspin wave resonance peak splitting and broadening by an-\ntidamping spin torque can be gained by neglecting the in-\ntrinsic nonlinearities \t nand\u0011nof the higher-order mode\njni. Setting \t n= 0 and\u0011n= 0 in Eqs. (3) and (4) allows\nus to reduce the equation of motion for the quasi-uniform\nmode amplitudej\u0016ajto the standard equation for a single-\nmode damped driven oscillator (Supplemental Material)\nwhere a constant damping parameter \u000b0is replaced by\nan e\u000bective frequency-dependent nonlinear damping pa-\nrameter\u000be\u000b\n0:\n\u000be\u000b\n0=\u000b0+\u0014\n\u00110+4\u000bn 2\nn\n(2!\u0000!n)2+ 4\u000b2n!2\u0015\nj\u0016aj2(5)and the resonance frequency is replaced by an e\u000bective\nresonance frequency:\n!e\u000b\n0=!0+\u0014\n\t0+2j nj2(2!\u0000!n)\n(2!\u0000!n)2+ 4\u000b2n!2\u0015\nj\u0016aj2(6)\nEquation (5) clearly shows that the damping parame-\nter of the quasi-uniform mode itself becomes a resonant\nfunction of the drive frequency with a maximum at half\nthe frequency of the higher order mode ( !=1\n2!n). The\namplitude and the width of this resonance in \u000be\u000b\n0(!) are\ndetermined by the intrinsic damping parameter \u000bnof\nthe higher-order mode jni. If\u000bnis su\u000eciently small,\nthe quasi-uniform mode damping is strongly enhanced\nat!=1\n2!n, which leads to a decrease of the quasi-\nuniform mode amplitude at this drive frequency. If the\ndrive frequency is shifted away from1\n2!nto either higher\nor lower values, the damping decreases, which can re-\nsult in an increase of the quasi-uniform mode amplitude\nj\u0016aj. Therefore, the amplitude of the quasi-uniform mode\nj\u0016aj(!) can exhibit a local minimum at !=1\n2!n. Due to\nits nonlinear origin, the tendency to form a local min-\nimum inj\u0016aj(!) at1\n2!nis enhanced with increasing j\u0016aj.\nSincej\u0016ajis large near the resonance frequency !0, tun-\ning!0to be equal to1\n2!ngreatly ampli\fes the e\u000bect of\nlocal minimum formation in j\u0016aj(!). This qualitative ar-\ngument based on Equation (5) explains the data in Fig. 2\n{ the split-peak nonlinear resonance of the quasi-uniform\nmode is only observed when external magnetic \feld tunes\nthe spin wave eigenmode frequencies to the three-magnon\nresonance condition !0=1\n2!n.\nEquation (6) reveals that the nonlinear frequency shift\nof the quasi-uniform mode is also a resonant function of\nthe drive frequency. In contrast to the nonlinear damping\nresonance described by Equation (5), the frequency shift\nresonance is an antisymmetric function of !\u00001\n2!n. The\nnonlinear shift is negative for ! <1\n2!nand thus causes\na fold-over towards lower frequencies while it is positive\nfor!>1\n2!ncausing fold-over towards higher frequencies.\nAt the center of the resonance pro\fle, the three-magnon\nprocess induces no frequency shift. This double-sided\nfold-over also contributes to the formation of the split-\npeak lineshape of the resonance shown in Figs. 2(c) and\n2(d) and to the linewidth broadening. As with the non-\nlinear damping resonance, the antisymmetric nonlinear\nfrequency shift and the double-sided fold-over become\ngreatly ampli\fed when the spin wave mode frequencies\nare tuned near the three-magnon resonance !0=1\n2!n.\nEquations (5) and (6) also shed light on the origin\nof the quasi-uniform mode line broadening by the an-\ntidamping spin torque. The antidamping spin torque in-\ncreases the quasi-uniform mode amplitude j\u0016ajvia transfer\nof angular momentum from spin current to the mode [57].\nSince the nonlinear damping and the nonlinear frequency\nshift are both proportional to j\u0016aj2and both contribute to\nthe line broadening, the antidamping spin torque can in-\ndeed give rise to the line broadening. Equation (5) reveals\ntwo competing e\u000bects of the antidamping spin torque on\nthe quasi-uniform mode damping parameter \u000be\u000b\n0: spin6\ntorque from Idcdecreases the linear component of the\ndamping parameter \u000b0!\u000b0(1\u0000Idc=Ij0i\nc) and increases\nthe nonlinear component via increased j\u0016aj2. Whether the\nantidamping spin torque decreases or increases the spec-\ntral linewidth of the mode depends on the system param-\neters. Our numerical solution of Eqs. (3) and (4) shown\nin Fig. 3 clearly demonstrates that the antidamping spin\ntorque can strongly increase the linewidth of the quasi-\nuniform mode when the three-magnon resonance condi-\ntion!0=1\n2!nis satis\fed. Furthermore, we \fnd that\nthe three-magnon process exhibits no threshold behav-\nior upon increasing amplitude (Supplemental Material)\nor decreasing intrinsic damping.\nThe key requirement for observation of the resonant\nnonlinear damping is the discreteness of the magnon\nspectrum imposed by geometric con\fnement in the\nnanoscale ferromagnet. The split-peak nonlinear reso-\nnance discovered in this work cannot be realized in bulk\nferromagnets because the three-magnon resonance con-\ndition in bulk is not only valid at the uniform mode\nfrequency!0=1\n2!nbut instead in a broad frequency\nrange. Owing to the magnon spectrum continuity in\nbulk, shifting the excitation frequency away from !0does\nnot suppress the three-magnon scattering of the uniform\nmode { it simply shifts it from one group of magnons to\nanother [29, 40]. Therefore, the amplitude of the uni-\nform mode does not increase when the drive frequency is\nshifted away from !0and the split-peak resonance is not\nrealized.\nWe expect that the resonant nonlinear damping dis-\ncovered in this work will have strong impact on the\nperformance of spin torque devices such as spin torque\nmagnetic memory, spin torque nanooscillators and spin\ntorque microwave detectors. Since all these devices rely\non large-amplitude oscillations of magnetization driven\nby spin torque, the amplitude limiting resulting from the\nresonant nonlinear damping is expected to have detri-\nmental e\u000bect on the device performance.V. CONCLUSIONS\nIn conclusion, our measurements demonstrate that\nmagnetic damping of spin wave modes in a nanoscale\nferromagnet has a strong nonlinear component of reso-\nnant character that appears at a discrete set of magnetic\n\felds corresponding to resonant three-magnon scattering.\nThis strong resonant nonlinearity can give rise to unusual\nspin wave resonance pro\fle with a local minimum at the\nresonance frequency in sharp contrast to the properties\nof the linear and nonlinear spin wave resonances in bulk\nferromagnets. The resonant nonlinearity has a profound\ne\u000bect on the response of the nanomagnet to spin torque.\nAntidamping spin torque, that reduces the quasi-uniform\nspin wave mode damping at magnetic \felds far from the\nresonant three-magnon regime, can strongly enhance the\ndamping in the resonant regime. This inversion of the\ne\u000bect of spin torque on magnetization dynamics by the\nresonant nonlinearity is expected to have signi\fcant im-\npact on the performance of nanoscale spin torque devices\nsuch as magnetic memory and spin torque oscillators.\nACKNOWLEDGMENTS\nThis work was supported by the National Science\nFoundation through Grants No. DMR-1610146, No.\nEFMA-1641989 and No. ECCS-1708885. We also ac-\nknowledge support by the Army Research O\u000ece through\nGrant No. W911NF-16-1-0472 and Defense Threat Re-\nduction Agency through Grant No. HDTRA1-16-1-0025.\nA. M. G. thanks CAPES Foundation, Ministry of Educa-\ntion of Brazil for \fnancial support. R.E.A acknowledges\nFinanciamiento Basal para Centros Cienti\fcos y Tec-\nnologicos de Excelencia under project FB 0807 (Chile),\nand Grant ICM P10-061-F by Fondo de Innovacion para\nla Competitividad-MINECON. B.A.I. was supported by\nthe National Academy of Sciences of Ukraine via project\n#1/17-N and by the Program of NUST \"MISiS\" (grant\nNo. K2-2017-005), implemented by a governmental de-\ncree dated 16th of March 2013, No. 211.\n[1] Luqiao Liu, Chi-Feng Pai, Y. Li, H. W. Tseng, D. C.\nRalph, and R. A. Buhrman, \\Spin-torque switching with\nthe giant spin Hall e\u000bect of tantalum,\" Science 336, 555{\n558 (2012).\n[2] S. Il. Kiselev, J.C. Sankey, I.N. Krivorotov, N.C. Emley,\nR.J. Schoelkopf, R.A. Buhrman, and D.C. Ralph, \\Mi-\ncrowave oscillations of a nanomagnet driven by a spin-\npolarized current,\" Nature 425, 380{383 (2003).\n[3] William H. Rippard, Alina M. Deac, Matthew R. Pufall,\nJustin M. Shaw, Mark W. Keller, Stephen E. Russek,\nGerrit E. W. Bauer, and Claudio Serpico, \\Spin-transfer\ndynamics in spin valves with out-of-plane magnetized\nCoNi free layers,\" Phys. Rev. B 81, 014426 (2010).[4] D. Houssameddine, U. Ebels, B. Delat, B. Rodmacq,\nI. Firastrau, F. Ponthenier, M. Brunet, C. Thirion, J.-\nP. Michel, L. Prejbeanu-Buda, M.-C. Cyrille, O. Redon,\nand B. Dieny, \\Spin-torque oscillator using a perpendic-\nular polarizer and a planar free layer,\" Nat. Mater. 6,\n447{453 (2007).\n[5] A. Houshang, E. Iacocca, P. Drrenfeld, S. R. Sani, J. ker-\nman, and R. K. Dumas, \\Spin-wave-beam driven syn-\nchronization of nanocontact spin-torque oscillators,\" Nat.\nNanotech. 11, 280{286 (2016).\n[6] Vladislav E. Demidov, Sergei Urazhdin, Henning Ul-\nrichs, Vasyl Tiberkevich, Andrei Slavin, Dietmar Baither,\nGuido Schmitz, and Sergej O. Demokritov, \\Magnetic\nnano-oscillator driven by pure spin current,\" Nat. Mater.7\n11, 1028{1031 (2012).\n[7] Ferran Maci\u0012 a, Dirk Backes, and Andrew D. Kent, \\Sta-\nble magnetic droplet solitons in spin-transfer nanocon-\ntacts,\" Nat. Nanotech. 9, 992{996 (2014).\n[8] Zhiliang Cheng, Ajlan Al Zaki, James Z. Hui, Vladimir R.\nMuzykantov, and Andrew Tsourkas, \\Multifunctional\nnanoparticles: Cost versus bene\ft of adding targeting\nand imaging capabilities,\" Science 338, 903{910 (2012).\n[9] Edward Kai-Hua Chow and Dean Ho, \\Cancer\nnanomedicine: From drug delivery to imaging,\" Science\nTransl. Med. 5, 216rv4 (2013).\n[10] Miriam Colombo, Susana Carregal-Romero, Maria F.\nCasula, Lucia Gutierrez, Maria P. Morales, Ingrid B.\nBohm, Johannes T. Heverhagen, Davide Prosperi, and\nWolfgang. J. Parak, \\Biological applications of magnetic\nnanoparticles,\" Chem. Soc. Rev. 41, 4306{4334 (2012).\n[11] Michael Bietenbeck, Anca Florian, Cornelius Faber, Udo\nSechtem, and Ali Yilmaz, \\Remote magnetic targeting\nof iron oxide nanoparticles for cardiovascular diagnosis\nand therapeutic drug delivery: where are we now?\" Intl.\nJ. Nanomed. 11, 3191 (2016).\n[12] Huilin Shao, Tae-Jong Yoon, Monty Liong, Ralph\nWeissleder, and Hakho Lee, \\Magnetic nanoparticles\nfor biomedical nmr-based diagnostics,\" Beilstein J. Nan-\notech. 1, 142 (2010).\n[13] P.W. Goodwill, A. Tamrazian, L.R. Croft, C.D. Lu, E.M.\nJohnson, R. Pidaparthi, R.M. Ferguson, A.P. Khandhar,\nK.M. Krishnan, and S.M. Conolly, \\Ferrohydrodynamic\nrelaxometry for magnetic particle imaging,\" Appl. Phys.\nLett. 98, 262502 (2011).\n[14] Patrick William Goodwill, Emine Ulku Saritas,\nLaura Rose Croft, Tyson N. Kim, Kannan M. Kr-\nishnan, David V. Scha\u000ber, and Steven M. Conolly,\n\\X-space mpi: Magnetic nanoparticles for safe medical\nimaging,\" Adv. Mater. 24, 3870{3877 (2012).\n[15] R. Matthew Ferguson, Kevin R. Minard, Amit P. Khand-\nhar, and Kannan M. Krishnan, \\Optimizing magnetite\nnanoparticles for mass sensitivity in magnetic particle\nimaging,\" Med. Phys. 38, 1619{1626 (2011).\n[16] Andrei Slavin and Vasil Tiberkevich, \\Nonlinear auto-\noscillator theory of microwave generation by spin-\npolarized current,\" IEEE Trans. Magn. 45, 1875{1918\n(2009).\n[17] Jian Zhu, J. A. Katine, Graham E. Rowlands, Yu-Jin\nChen, Zheng Duan, Juan G. Alzate, Pramey Upad-\nhyaya, Juergen Langer, Pedram Khalili Amiri, Kang L.\nWang, and Ilya N. Krivorotov, \\Voltage-induced ferro-\nmagnetic resonance in magnetic tunnel junctions,\" Phys.\nRev. Lett. 108, 197203 (2012).\n[18] S. Miwa, S. Ishibashi, H. Tomita, T. Nozaki, E. Tamura,\nK. Ando, N. Mizuochi, T. Saruya, H. Kubota,\nK. Yakushiji, T. Taniguchi, H. Imamura, A. Fukushima,\nS. Yuasa, and Y. Suzuki, \\Highly sensitive nanoscale\nspin-torque diode,\" Nat. Mater. 13, 50{56 (2014).\n[19] Bin Fang, Mario Carpentieri, Xiaojie Hao, Hongwen\nJiang, Jordan A. Katine, Ilya N. Krivorotov, Berthold\nOcker, Juergen Langer, Kang L. Wang, Baoshun Zhang,\nBruno Azzerboni, Pedram Khalili Amiri, Giovanni Finoc-\nchio, and Zhongming Zeng, \\Giant spin-torque diode\nsensitivity in the absence of bias magnetic \feld,\" Nat.\nCommun. 7, 11259 (2016).\n[20] A. M. Gon\u0018 calves, I. Barsukov, Y.-J. Chen, L. Yang, J. A.\nKatine, and I. N. Krivorotov, \\Spin torque ferromagnetic\nresonance with magnetic \feld modulation,\" Appl. Phys.Lett. 103, 172406 (2013).\n[21] A.A. Tulapurkar, Y. Suzuki, A. Fukushima, H. Kub-\nota, H. Maehara, K. Tsunekawa, D.D. Djayaprawira,\nN. Watanabe, and S. Yuasa, \\Spin-torque diode e\u000bect in\nmagnetic tunnel junctions,\" Nature 438, 339{342 (2005).\n[22] J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N. Kriv-\norotov, R. A. Buhrman, and D. C. Ralph, \\Spin-transfer-\ndriven ferromagnetic resonance of individual nanomag-\nnets,\" Phys. Rev. Lett. 96, 227601 (2006).\n[23] Motoya Shinozaki, Eriko Hirayama, Shun Kanai, Hideo\nSato, Fumihiro Matsukura, and Hideo Ohno, \\Damp-\ning constant in a free layer in nanoscale CoFeB/MgO\nmagnetic tunnel junctions investigated by homodyne-\ndetected ferromagnetic resonance,\" Appl. Phys. Exp. 10,\n013001 (2017).\n[24] Christopher J. Safranski, Yu-Jin Chen, Ilya N. Krivo-\nrotov, and Jonathan Z. Sun, \\Material parameters of\nperpendicularly magnetized tunnel junctions from spin\ntorque ferromagnetic resonance techniques,\" Appl. Phys.\nLett. 109, 132408 (2016).\n[25] Michael Harder, Yongsheng Gui, and Can-Ming Hu,\n\\Electrical detection of magnetization dynamics via spin\nrecti\fcation e\u000bects,\" Phys. Rep. 661, 1{59 (2016).\n[26] O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer,\nS. D. Bader, and A. Ho\u000bmann, \\Quantifying spin Hall\nangles from spin pumping: Experiments and theory,\"\nPhys. Rev. Lett. 104, 046601 (2010).\n[27] C. Liu, Y. Boyko, C. C. Geppert, K. D. Christie,\nG. Stecklein, S. J. Patel, C. J. Palmstrm, and P. A.\nCrowell, \\Electrical detection of ferromagnetic resonance\nin ferromagnet/n-GaAs heterostructures by tunneling\nanisotropic magnetoresistance,\" Appl. Phys. Lett. 105,\n212401 (2014).\n[28] Shinji Miwa, Junji Fujimoto, Philipp Risius, Kohei\nNawaoka, Minori Goto, and Yoshishige Suzuki, \\Strong\nbias e\u000bect on voltage-driven torque at epitaxial Fe-MgO\ninterface,\" Phys. Rev. X 7, 031018 (2017).\n[29] C\u0013 esar L. Ord\u0013 o~ nez-Romero, Boris A. Kalinikos, Pavol\nKrivosik, Wei Tong, Pavel Kabos, and Carl E. Pat-\nton, \\Three-magnon splitting and con\ruence processes\nfor spin-wave excitations in yttrium iron garnet \flms:\nWave vector selective Brillouin light scattering measure-\nments and analysis,\" Phys. Rev. B 79, 144428 (2009).\n[30] H. Schultheiss, X. Janssens, M. van Kampen, F. Ciub-\notaru, S. J. Hermsdoerfer, B. Obry, A. Laraoui, A. A.\nSerga, L. Lagae, A. N. Slavin, B. Leven, and B. Hille-\nbrands, \\Direct current control of three magnon scat-\ntering processes in spin-valve nanocontacts,\" Phys. Rev.\nLett. 103, 157202 (2009).\n[31] C. T. Boone, J. A. Katine, J. R. Childress, V. Tiberke-\nvich, A. Slavin, J. Zhu, X. Cheng, and I. N. Krivorotov,\n\\Resonant nonlinear damping of quantized spin waves\nin ferromagnetic nanowires: A spin torque ferromagnetic\nresonance study,\" Phys. Rev. Lett. 103, 167601 (2009).\n[32] Hidekazu Kurebayashi, Oleksandr Dzyapko, Vladislav E.\nDemidov, Dong Fang, Andrew J. Ferguson, and\nSergej O. Demokritov, \\Controlled enhancement of spin-\ncurrent emission by three-magnon splitting,\" Nat. Mater.\n10, 660{664 (2011).\n[33] R. N. Costa Filho, M. G. Cottam, and G. A. Farias,\n\\Microscopic theory of dipole-exchange spin waves in fer-\nromagnetic \flms: Linear and nonlinear processes,\" Phys.\nRev. B 62, 6545{6560 (2000).8\n[34] Alina M Deac, Akio Fukushima, Hitoshi Kubota, Hi-\nroki Maehara, Yoshishige Suzuki, Shinji Yuasa, Yoshinori\nNagamine, Koji Tsunekawa, David D Djayaprawira, and\nNaoki Watanabe, \\Bias-driven high-power microwave\nemission from MgO-based tunnel magnetoresistance de-\nvices,\" Nat. Phys. 4, 803{809 (2008).\n[35] V. E. Demidov, S. Urazhdin, E. R. J. Edwards, M. D.\nStiles, R. D. McMichael, and S. O. Demokritov, \\Con-\ntrol of magnetic \ructuations by spin current,\" Phys. Rev.\nLett. 107, 107204 (2011).\n[36] G. D. Fuchs, J. C. Sankey, V. S. Pribiag, L. Qian,\nP. M. Braganca, A. G. F. Garcia, E. M. Ryan, Zhi-\nPan Li, O. Ozatay, D. C. Ralph, and R. A. Buhrman,\n\\Spin-torque ferromagnetic resonance measurements of\ndamping in nanomagnets,\" Appl. Phys. Lett. 91, 062507\n(2007).\n[37] G.A. Melkov, D.V. Slobodianiuk, V.S. Tiberkevich,\nG. de Loubens, O. Klein, and A.N. Slavin, \\Nonlinear\nferromagnetic resonance in nanostructures having dis-\ncrete spectrum of spin-wave modes,\" IEEE Magn. Lett.\n4, 4000504 (2013).\n[38] M. Helsen, A. Gangwar, J. De Clercq, A. Vansteenkiste,\nM. Weigand, C. H. Back, and B. Van Waeyenberge,\n\\Non-linear radial spinwave modes in thin magnetic\ndisks,\" Appl. Phys. Lett. 106, 032405 (2015).\n[39] Jan Podbielski, Detlef Heitmann, and Dirk Grundler,\n\\Microwave-assisted switching of microscopic rings: Cor-\nrelation between nonlinear spin dynamics and critical mi-\ncrowave \felds,\" Phys. Rev. Lett. 99, 207202 (2007).\n[40] H. Suhl, \\The theory of ferromagnetic resonance at high\nsignal powers,\" J. Phys. Chem. Solids 1, 209{227 (1957).\n[41] Hans G Bauer, Peter Majchrak, Torsten Kachel, Chris-\ntian H Back, and Georg Woltersdorf, \\Nonlinear spin-\nwave excitations at low magnetic bias \felds,\" Nat. Com-\nmun. 6, 8274 (2015).\n[42] M. Bauer, O. B uttner, S. O. Demokritov, B. Hillebrands,\nV. Grimalsky, Yu. Rapoport, and A. N. Slavin, \\Obser-\nvation of spatiotemporal self-focusing of spin waves in\nmagnetic \flms,\" Phys. Rev. Lett. 81, 3769{3772 (1998).\n[43] Arnold Markovich Kosevich, B.A. Ivanov, and A.S.\nKovalev, \\Magnetic solitons,\" Phys. Rep. 194, 117{238\n(1990).\n[44] A. N. Slavin and I. V. Rojdestvenski, \\Bright and dark\nspin wave envelope solitons in magnetic \flms,\" IEEE\nTrans. Magn. 30, 37{45 (1994).\n[45] Mingzhong Wu, Boris A. Kalinikos, and Carl E. Patton,\n\\Self-generation of chaotic solitary spin wave pulses in\nmagnetic \flm active feedback rings,\" Phys. Rev. Lett.\n95, 237202 (2005).\n[46] V. V. Naletov, G. de Loubens, V. Charbois, O. Klein,\nV. S. Tiberkevich, and A. N. Slavin, \\Ferromagnetic\nresonance spectroscopy of parametric magnons excited\nby a four-wave process,\" Phys. Rev. B 75, 140405 (2007).\n[47] Y. Khivintsev, Bijoy Kuanr, T. J. Fal, M. Haftel, R. E.\nCamley, Z. Celinski, and D. L. Mills, \\Nonlinear fer-\nromagnetic resonance in permalloy \flms: A nonmono-\ntonic power-dependent frequency shift,\" Phys. Rev. B 81,\n054436 (2010).\n[48] Bivas Rana, Yasuhiro Fukuma, Katsuya Miura, Hiro-\nmasa Takahashi, and YoshiChika Otani, \\E\u000bect of ex-\ncitation power on voltage induced local magnetization\ndynamics in an ultrathin CoFeB \flm,\" Sci. Rep. 7, 2318\n(2017).[49] M. d'Aquino, A. Quercia, V. Scalera, S. Perna,\nG. Bertotti, I. D. Mayergoyz, and C. Serpico, \\Ana-\nlytical treatment of nonlinear ferromagnetic resonance in\nnanomagnets,\" IEEE Trans. Magn. 53, 4301005 (2017).\n[50] I. N. Krivorotov, N. C. Emley, R. A. Buhrman, and D. C.\nRalph, \\Time-domain studies of very-large-angle mag-\nnetization dynamics excited by spin transfer torques,\"\nPhys. Rev. B 77, 054440 (2008).\n[51] Feng Guo, Lyubov M. Belova, and Robert D. McMichael,\n\\Nonlinear ferromagnetic resonance shift in submicron\npermalloy ellipses,\" Phys. Rev. B 91, 064426 (2015).\n[52] Hans T. Nembach, Justin M. Shaw, Carl T. Boone, and\nT. J. Silva, \\Mode- and size-dependent Landau-Lifshitz\ndamping in magnetic nanostructures: Evidence for non-\nlocal damping,\" Phys. Rev. Lett. 110, 117201 (2013).\n[53] Y. Li and W.E. Bailey, \\Wave-number-dependent Gilbert\ndamping in metallic ferromagnets,\" Phys. Rev. Lett. 116,\n117602 (2016).\n[54] Koji Sekiguchi, Keisuke Yamada, Soo-Man Seo, Kyung-\nJin Lee, Daichi Chiba, Kensuke Kobayashi, and Teruo\nOno, \\Time-domain measurement of current-induced\nspin wave dynamics,\" Phys. Rev. Lett. 108, 017203\n(2012).\n[55] V. Lauer, D. A. Bozhko, T. Brcher, P. Pirro, V. I.\nVasyuchka, A. A. Serga, M. B. Jung\reisch, M. Agrawal,\nYu. V. Kobljanskyj, G. A. Melkov, C. Dubs, B. Hille-\nbrands, and A. V. Chumak, \\Spin-transfer torque based\ndamping control of parametrically excited spin waves in\na magnetic insulator,\" Appl. Phys. Lett. 108, 012402\n(2016).\n[56] Chi Zhang, Yong Pu, Sergei A. Manuilov, Shane P.\nWhite, Michael R. Page, Erick C. Blomberg, Denis V.\nPelekhov, and P. C. Hammel, \\Engineering the spec-\ntrum of dipole \feld-localized spin-wave modes to enable\nspin-torque antidamping,\" Phys. Rev. Appl. 7, 054019\n(2017).\n[57] S. M. Rezende, F. M. de Aguiar, and A. Azevedo,\n\\Magnon excitation by spin-polarized direct currents\nin magnetic nanostructures,\" Phys. Rev. B 73, 094402\n(2006).Supplemental Material:\nGiant resonant nonlinear damping in nanoscale ferromagnets\nI. Barsukov,1H. K. Lee,1A. A. Jara,1Y.-J. Chen,1A. M. Gon¸ calves,1C.\nSha,1J. A. Katine,2R. E. Arias,3B. A. Ivanov,4, 5and I. N. Krivorotov1\n1Physics and Astronomy, University of California, Irvine, CA 92697, USA\n2Western Digital, 5600 Great Oaks Parkway, San Jose, CA 95119, USA\n3Departamento de F´ ısica, CEDENNA, FCFM, Universidad de Chile, Santiago, Chile\n4Institute of Magnetism, National Academy of Sciences of Ukraine, Vernadsky av. 36 B, Kyiv, 03142, Ukraine\n5National University of Science and Technology MISiS, Moscow, 119049, Russian Federation\nI. METHODS\nA. Linewidth evaluation\nAll measurements presented were carried out with magnetic field applied along the easy axis of the MTJ devices so\nthat the magnetic moments of the free and pinned layers are collinear to each other. In this geometry, the ST-FMR sig-\nnals are dominated by photo-resistance contribution and are proportional to the square of the transverse component of\nthe dynamic magnetization magnetization [1], which allows us to directly compare calculated |a|2(ω) resonance curves\nto measured ST-FMR resonance curves ˜Vmix(f) and toVmix(f) approximated by numerical integration/integraltext˜Vmix(f)df.\nWhenVmix(f) and|a|2(ω) are single-peak curves, they are fit to a sum of symmetric and antisymmetric Lorentzian\ncurves with identical central frequencies and linewidth parameters as described in Ref. [2], and the spectral linewidth\nis determined as half-width at the half-maximum of the symmetric Lorentzian curve.\nIn order to quantify the linewidth of the split-peak resonance profile, we introduce a fitting function that is a sum\nof two Lorentzian curves with different central frequencies separated by δf. The half width of the resonance profile\n∆f0is then defined as the average of the half widths of the two Lorentzians plus δf/2.\nSupplemental Figure 1. Spatial profiles of spin wave eigenmodes. Normalized amplitude and phase of the three lowest frequency\nspin wave eigenmodes of the MTJ free layer, given by micromagnetic simulations.\nB. Micromagnetic simulations\nMicromagnetic simulations were performed using OOMMF software [3, 4]. To account for all magnetic interactions\nin the MTJ, a three dimensional model was employed with three ferromagnetic layers: free, SAF top and SAF bottom.\nWe use material parameters obtained from the measurements and/or their accepted literature values (see Ref. [2] for2\nthe MTJ structure and fabrication details). Magnetization dynamics is excited by a combined pulse of spin torque\nand Oersted field, resulting from a sinc-shaped spatially uniform current pulse. The spatial profile of the Oersted\nfield corresponds to that of a long wire with elliptical cross section. The direction of the spin torque vector acting\non the free layer is determined by the magnetization orientation of the SAF top layer. The spectrum of spin wave\neigenmodes is obtained via fast Fourier transform (FFT) of the time dependent components of the layers’ magnetic\nmoment. Spatial mapping of the resulting Fourier amplitude and phase at a given frequency provides the mode\nprofiles (Supplmental Fig. 1). The observed excitations are confirmed to be spin wave modes localized to the free\nlayer. SAF modes are found at much higher frequencies than the free layer modes, and their frequencies are found to\nbe incommensurable to the free layer quasi-uniform mode frequency [5].\nII. SOLUTION OF THE EQUATIONS OF MOTION\nThe Hamiltonian equations of motion describing the coupled dissipative dynamics of the quasi-uniform ( a) and the\nhigher-order ( b) spin wave modes are:\nida\ndt=∂H\n∂a†+∂Q\n∂(da†/dt)(1)\nidb\ndt=∂H\n∂b†+∂Q\n∂(db†/dt)(2)\nwhereHis the Hamiltonian of the system and Qis the dissipation function, given by:\nH=ω0a†a+ωnb†b+1\n2Ψ0a†a†aa+1\n2Ψnb†b†bb+ (ψ∗\nnaab†+ψna†a†b) +ζ{exp(−iωt)a†+ exp(iωt)a} (3)\nQ=da†\ndtda\ndt(α0+η0a†a) +db†\ndtdb\ndt(αn+ηnb†b) (4)\nBy using Eq. (3) and Eq. (4) in Eq. (1) and Eq. (2), the Hamiltonian equations can be written as:\nida\ndt−(α0+η0a†a)da\ndt=ω0a+ 2ψna†b+ Ψ 0a†aa+ζexp(−iωt) (5)\nidb\ndt−(αn+ηnb†b)db\ndt=ωnb+ψ∗\nnaa+ Ψnb†bb (6)\nUsing a periodic ansatz a= ¯aexp(−iωt) andb=¯bexp(−2iωt) in Eq. (5) and Eq. (6), where ¯ aand¯bare complex\namplitudes, reduces the Hamiltonian equations to a set of two algebraic equation for the complex amplitudes:\n/parenleftbig\nω−ω0−Ψ0|¯a|2+i(α0+η0|¯a|2)ω/parenrightbig\n¯a−2ψn¯a∗¯b=ζ (7)\n/parenleftbig\n2ω−ωn−Ψn|¯b|2+ 2i(αn+ηn|¯b|2)ω/parenrightbig¯b=ψ∗\nn¯a2(8)\nWe solve Eq. (8) for ¯band multiply the numerator and denominator of this expression by the complex conjugate of\nthe denominator:\n¯b=ψ∗\nn¯a2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n−i2(αn+ηn|¯b|2)ω\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(9)\nthen we multiply Eq. (9) by2ψn¯a∗\n¯aand evaluate the real and imaginary parts.\n/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=|ψn|2|¯a|22/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(10)\n/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=|ψn|2|¯a|2−4(αn+ηn|¯b|2)ω\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(11)\nBy taking the modulus of Eq. (8), we obtain:\n|¯a|2=|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2 (12)3\nUsing Eq. (12) in Eqs. (10-11), we derive:\n/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(13)\n/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=−4(αn+ηn|¯b|2)ω|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(14)\nTaking the modulus squared of Eq. (7):\n/braceleftBigg/parenleftbigg\nω−ω0−Ψ0|¯a|2−/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg/parenrightbigg2\n+/parenleftbigg\n(α0+η0|¯a|2)ω−/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg/parenrightbigg2/bracerightBigg\n|¯a|2=ζ2(15)\nand using Equations (12)–(14) in Eq. (15) gives us an algebraic equation for the absolute value of the higher order\nmode amplitude|¯b|:\n\n\n\nω−ω0−Ψ0|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2−2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2\n2\n+\n\n/parenleftbigg\nα0+η0|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2/parenrightbigg\nω−−4(αn+ηn|¯b|2)ω|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2\n2\n\n×\n|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2=ζ2\n(16)\nAfter numerically solving Eq. (16) for |¯b|, and using it in Eq. (12), we can calculate the amplitude of the quasi-uniform\nmode|¯a|.\nIII. EFFECTS OF THE DRIVE AMPLITUDE AND INTRINSIC NONLINEARITITES\nTo understand the impact of the intrinsic nonlinearity parameters (Ψ 0, Ψn,η0,ηn) on the quasi-uniform spin wave\nmode resonance, we plot the numerical solution of Eq. (16) in Supplemental Figure 2. Each panel of this figure shows\na reference lineshape of the resonance calculated with all intrinsic nonlinearity parameters set to zero (red curve) and\na lineshape calculated with one of the intrinsic nonlinearity parameter different from zero (blue curve). This figure\nreveals that increasing η0decreases the mode amplitude and slightly increases the linewidth. Increasing ηndecreases\nthe degree of the double-peak lineshape splitting. Increasing Ψ nincreases the lineshape asymmetry. Increasing Ψ 0\nincreases lineshape asymmetry and induces fold-over.\nSupplemental Figure 3 shows the linewidth as a function of the drive amplitude for three scenarios, where the\nintrinsic nonlinearities Ψ 0, Ψn,η0,ηnare set to zero for simplicity. If the coupling parameter is zero, ψn= 0, the\nlinewidth does not depend on the drive amplitude, as expected for a single-mode linear oscillator. The second case\ndemonstrates that the linewidth remains constant when the product ψn·ζis constant. For a constant non-zero\ncoupling parameter, the linewidth shows an increase with the drive amplitude. This observation allows us to employ a\nsingle fitting parameter ( ψn·ζ) to fit the data in Fig. 1b. This conjecture can be confirmed analytically by introducing\na normalized spin wave amplitude ˆ a=ψn¯a, which allows us to rewrite Eq. (16) omitting all intrinsic nonlinearities\ninto the following form:\nω/bracketleftbigg\n1 +iα0+i4αn|ˆa|2\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\nˆa−ω0ˆa−2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2|ˆa|2ˆa=ψnζ (17)\nThis equation describes an effective single-mode nonlinear oscillator with renormalized excitation amplitude ψnζ.4\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nη0\n2 2.5 3 3.500.010.02\nFrequency (GHz)|a|20\nηn\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nΨ0\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nΨn|a|2|a|2|a|2a b\nc d\nSupplemental Figure 2. Effect of intrinsic nonlinearities on the quasi-uniform spin wave resonance lineshape. Spectral lineshape\nof the quasi-uniform spin wave mode resonance |¯a|2(ω) at the three-magnon resonance condition 2 ω0=ωncalculated by\nnumerically solving Eq. (16). The red curve is a reference lineshape calculated with all intrinsic nonlinearity parameters\n(η0,ηn,Ψ0,Ψn) set to zero. The blue lineshape in each panel is calculated with one of the intrinsic nonlinearity parameters set\nto a non-zero value: (a) η0= 1.325·10−24J, (b)ηn= 3.313·10−24J, (c) Ψ 0= 1.325·10−24J, (d) Ψ n= 1.325·10−23J. Other\nparameters employed in the calculation are: ω0= 2π·2.63 GHz,ωn= 2π·5.26 GHz;α0= 0.02662,αn= 0.03042 atIdc= 0;\nψn·ζ=h2·0.006 GHz2, wherehis the Planck constant.\n0 0.05 0.1 0.15 0.20.150.20.25\nζ h-1 (GHz)∆f0 (GHz)\n(i)(ii)(iii)\nSupplemental Figure 3. Effect of the drive amplitude on linewidth in the resonant three-magnon regime. Calculated linewidth\nof the quasi-uniform spin wave mode as a function of the drive amplitude ζfor different values of the mode coupling parameter\nψn. (i) Green: ψn= 0, (ii) red: variable ψnwith a constraint ψn·ζ=h2·0.006 GHz2, and (iii) blue: ψn=h·0.1 GHz. All\nintrinsic nonlinearity parameters: Ψ 0, Ψn,η0andηnare set to zero. his the Planck constant. Other parameters employed in\nthe calculation are: ω0= 2π·2.63 GHz,ωn= 2π·5.26 GHz;α0= 0.02662 andαn= 0.03042 atIdc= 0.5\nIV. EFFECTIVE SINGLE-MODE NONLINEAR OSCILLATOR APPROXIMATION\nIf we neglect intrinsic nonlinearities Ψ nandηnof the higher order spin wave mode, Eq. (16) can be reduced to a\ncubic equation for ¯ aand solved analytically. This approximation allows us to obtain several important qualitative\ninsights into the properties of the resonant nonlinear damping of the quasi-uniform mode. By setting Ψ n= 0 and\nηn= 0 in Eq. (8), we obtain an exact solution for ¯b:\n¯b=ψ∗\nn¯a2\n2ω(1 +iαn)−ωn(18)\nUsing this result, we reduce Eq. (16) to a cubic algebraic equation for ¯ a:\nω/bracketleftbigg\n1 +i(α0+η0|¯a|2) +i4|ψn|2αn|¯a|2\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n¯a−ω0¯a−/bracketleftbigg\nΨ0+2|ψn|2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2¯a=ζ (19)\nThis equation describes the amplitude ¯ aof an effective single-mode nonlinear oscillator.\nIt is evident from Eq. (19) that the frequency of the quasi-uniform mode experiences a nonlinear shift:\nωeff\n0=ω0+/bracketleftbigg\nΨ0+2|ψn|2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2(20)\nThe nonlinear frequency shift has a well-pronounced antisymmetric resonant character near the resonance frequency\nωn/2, that arises from the resonant three-magnon scattering.\nFurther, it is clear from Eq. (19) that the effective damping of the quasi-uniform mode also acquires a term arising\nfrom the three-magnon interaction:\nαeff\n0=α0+/bracketleftbigg\nη0+4|ψn|2αn\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2(21)\nThe last term describes a resonant enhancement of the nonlinear damping by three-magnon scattering near the\nresonance frequency ωn/2. Strikingly, the magnitude of the resonant damping enhancement at ωn/2 increases when\nthe intrinsic damping of the higher order mode αndecreases. In the limit αn→0, the effective damping becomes\nαeff\n0→α0+/bracketleftbigg\nη0+2π|ψn|2\nωδ(2ω−ωn)/bracketrightbigg\n|¯a|2(22)\nwhereδis Dirac’s delta function. Equation (21) suggests that the effective damping of the quasi-uniform mode αeff\n0\ncan increase with increasing antidamping spin torque applied to the nanomagnet. Indeed, the antidamping spin torque\ntends to increase the amplitude [6] of the quasi-uniform mode |¯a|and decrease the intrinsic damping parameter of the\nhigher order mode αn→αn(1−Idc/I|n/angbracketright\nc), both enhancing the nonlinear damping term in Eq. (19). For a sufficiently\nlarge mode coupling parameter ψn, the enhancement of the nonlinear damping term by the antidamping spin torque\ncan exceed the reduction of the linear damping parameter α0→α0(1−Idc/I|0/angbracketright\nc) by the torque, leading to an increase\nofαeff\n0byIdc>0 and broadening of the quasi-uniform mode resonance by the antidamping spin torque. This scenario\nis indeed realized in the MTJ devices studied here as demonstrated by the data and calculations in Fig. 3.\nV. MODE COUPLING PARAMETER\nIn this Supplementary Note, we discuss how the coupling parameter between the spin wave modes, ψnin Eq. (3),\ncan be calculated. We consider a very thin, magnetically soft ferromagnetic disk with elliptical cross section, that\nis magnetized in-plane. Within a classical micromagnetic model, we include Zeeman, dipolar and exchange terms in\nthe free energy. An applied field Halong thexdirection (long axis of the ellipse) magnetizes the sample to a nearly\nuniform state. Through a classical Holstein-Primakoff transformation [7] we introduce variables c(/vector x,t) andc∗(/vector x,t) to\ndescribe the magnetization such that the magnetization magnitude is conserved:\nmx= 1−cc∗, m +=c√2−cc∗, m−=c∗√2−cc∗, (23)\nwhere/vector m=/vectorM/Ms, andm±≡mz±imy. Approximating the exchange energy to the fourth order in candc∗, the\nnormalized free energy of the disk, U≡E/4πM2\ns, is given by\nU/similarequal−hx/integraldisplay\n(1−cc∗) dV+ (lex)2/integraldisplay/bracketleftbigg\n/vector∇c·/vector∇c∗+1\n4c2(/vector∇c∗)2+1\n4c∗2(/vector∇c)2/bracketrightbigg\ndV−1\n2/integraldisplay\ndV/vectorhD(/vector m)·/vector m, (24)6\nwithhx≡H/4πMs,lex≡/radicalbig\nA/2πM2sis the exchange length, and /vectorhD(/vector m) =/vectorHD(/vector m)/4πMsis the normalized\ndemagnetizing field. The Landau-Lifshitz equations of motion in the new variables are: i˙c=δU/δc∗,i˙c∗=−δU/δc\nwitht/prime= 4πMs|γ|t.\nAssuming the normal modes involved in three magnon scattering dominate the magnetization dynamics, the free\nenergy in Eq. (24) can be written in terms of amplitudes of these modes, by expressing cin terms of aandb:\nc(/vector x,t)/similarequala(t)f(/vector x) +a∗(t)g(/vector x) +b(t)p(/vector x) +b∗(t)q(/vector x) (25)\nThe functions f,g,p,q can be determined from calculating the linear modes of oscillation of the sample. The terms of\nthe free energy proportional to aab∗anda∗a∗bdescribe the three-magnon process and the magnitude of these terms\ngives the coupling parameter ψn.\nIf the magnetization state is approximated as exactly uniform, the dipolar energy for a very thin film may be\napproximated as UD=m2\nz/2 = (c+c∗)2(1−cc∗/2), and in this case all three-magnon terms are zero. However, when\nthe effects due to the sample edges (such as spatial inhomogeneity of the demagnetization field and edge roughness)\nare taken into account, the equilibrium magnetization configuration is generally nonuniform. In this case, there are\nnon-zero three-magnon terms in the free energy expression. An explicit calculation of the corresponding overlap\nintegrals is necessary for a quantitative prediction of ψn. Refs. [8, 9] show such extensive calculations for circular disks\nand include explicit expressions for the exchange and dipolar energies.\n[1] Michael Harder, Yongsheng Gui, and Can-Ming Hu, “Electrical detection of magnetization dynamics via spin rectification\neffects,” Phys. Rep. 661, 1–59 (2016).\n[2] A. M. Gon¸ calves, I. Barsukov, Y.-J. Chen, L. Yang, J. A. Katine, and I. N. Krivorotov, “Spin torque ferromagnetic\nresonance with magnetic field modulation,” Appl. Phys. Lett. 103, 172406 (2013).\n[3] M. J. Donahue and D. G. Porter, OOMMF User’s Guide (National Institute of Standards and Technology, Gaithersburg,\nMD, 1999).\n[4] Robert D. McMichael and Mark D. Stiles, “Magnetic normal modes of nanoelements,” J. Appl. Phys. 97, 10J901 (2005).\n[5] P. S. Keatley, V. V. Kruglyak, A. Neudert, R. J. Hicken, V. D. Poimanov, J. R. Childress, and J. A. Katine, “Resonant\nenhancement of damping within the free layer of a microscale magnetic tunnel valve,” J. Appl. Phys. 117, 17B301 (2015).\n[6] J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N. Krivorotov, R. A. Buhrman, and D. C. Ralph, “Spin-transfer-driven\nferromagnetic resonance of individual nanomagnets,” Phys. Rev. Lett. 96, 227601 (2006).\n[7] T. Holstein and H. Primakoff, “Field dependence of the intrinsic domain magnetization of a ferromagnet,” Phys. Rev. 58,\n1098–1113 (1940).\n[8] D. Mancilla-Almonacid and R. E. Arias, “Instabilities of spin torque driven auto-oscillations of a ferromagnetic disk mag-\nnetized in plane,” Phys. Rev. B 93, 224416 (2016).\n[9] D. Mancilla-Almonacid and R. E. Arias, “Spin-wave modes in ferromagnetic nanodisks, their excitation via alternating\ncurrents and fields, and auto-oscillations,” Phys. Rev. B 95, 214424 (2017)." }, { "title": "1507.03308v1.Effective_90_degree_magnetization_rotation_in_Co2FeAl_thin_film_Piezoelectric_system_probed_by_microstripline_ferromagnetic.pdf", "content": "Effective 90-degree magnetization rotation in Co 2FeAl thin film/Piezoelectric system\nprobed by microstripline ferromagnetic resonance\nM. Gueye1, F. Zighem1,\u0003M. Belmeguenai1, M. Gabor2, C. Tiusan2, and D. Faurie1\n1LSPM-CNRS, Université Paris XIII, Sorbonne Paris Cité, 93430 Villetaneuse, France and\n2Center for Superconductivity, Spintronics and Surface Science, Technical University of Cluj-Napoca,\nStr. Memorandumului No. 28 RO-400114, Cluj-Napoca, Romania\n(Dated: July 7th2015)\nMicrostripline ferromagnetic resonance technique has been used to study the indirect magneto-\nelectric coupling occurring in an artificial magnetoelectric heterostructure consisting of a magne-\ntostrictive thin film cemented onto a piezoelectric actuator. Two different modes (sweep-field and\nsweep-frequency modes) of this technique have been employed to quantitatively probe the indirect\nmagnetoelectric coupling and to observe a voltage induced magnetization rotation (of 90 degree).\nThis latter has been validated by the experimental frequency variation of the uniform mode and by\nthe amplitude of the sweep-frequency spectra.\nIn recent decades, the development and the sophisti-\ncation of growth and characterization on the atomic and\nnanometric scale, has led to the elaboration of smart ma-\nterials which reveal captivating phenomena. In the midst\nof these materials, multiferroics magnetoelectric materi-\nals, currently at the cutting edge of spintronics, have at-\ntracted the attention of many groups [1–5]. A flurry of\nresearch activities has been launched in order to explore\nthe physics behind these materials. From the applica-\ntions standpoint, an alternative approach to overcome\nthe scientific impediments of weak magnetoelectric cou-\npling in single-phase multiferroics are two-phase systems\nofferromagneticandferroelectricconstituents[6]. Oneof\nthe major advantages of ferromagnetic/ferroelectric ex-\ntrinsic multiferroics as reported in the literature is the\nflexibility in the choice of the materials [6]. One of the\nmain idea behind the fabrication of artificial magneto-\nelectric multiferroics is the electric or voltage control of\nthe magnetization using electric fields for low power and\nultra fast next generation electronics [7, 8]. The straight-\nforward heterostructure allowing such a control appears\nto be the one presenting a piezoelectric/magnetostrictive\ninterface wherein the interaction vector is the voltage in-\nduced in-plane strains[6, 9–14].\nThe present study concerns the experimental observa-\ntion of the voltage induced magnetization rotation in an\nartificial magnetoelectric heterostructure by probing the\nmagnetic resonance of the uniform precession mode. A\nmethod based only on microstripline ferromagnetic res-\nonance (MS-FMR) experiments is presented [15]. The\ntwo different modes of this technique are presented and\nused to perform a quantitative characterization of the\neffective magnetoelectric coupling, as well as of the volt-\nage inducedmagnetization-rotation. Figures1-a) and-b)\nshow the artificial heterostructure composed of a 25 nm\nCo2FeAl(CFA)thinfilmgrownontoa125 µmthickpoly-\nimide (Kapton ®) flexible substrate and then cemented\non a piezoelectric actuator. The CFA film was deposited\n\u0003Electronic address: zighem@univ-paris13.fronto the Kapton ®substrate by rf sputtering. The de-\nposition residual pressure was of around 4\u000210\u00009mbar,\nwhile the working Ar pressure was of 1:3\u000210\u00003mbar. A\n5 nm thick Ta cap layer was deposited on the top of the\nCFA film in order to protect it from oxidation. CFA was\nchosen because of its non-negligible magnetostriction co-\nefficient at saturation even in a polycrystalline film with\nnopreferredorientation( \u0015\u001914\u000210\u00006)[16,17]whichwill\nproducestrongindirectmagnetoelectricfieldandalsobe-\ncause it is a serious candidate for spintronic applications\n[18–20]. Moreover, we have shown that Kapton substrate\ndoes not affect the whole system magnetoelectric behav-\nior because of its high compliance [9, 11].\nMS-FMR experiments can be carried out in two dif-\nferent modes: i) by fixing the microwave driving fre-\nquency (f0)and sweeping the applied magnetic field (H)\n[11, 15, 18] or ii) by fixing the applied magnetic field\n(H0)and sweeping the microwave frequency (f) [21].\nA resonance field ( Hresat fixedf0) and a frequency\nresonance ( fresat fixedH0) are extracted from these\ntwo modes each having their own advantage. Indeed,\nthe sweep-field mode is more sensitive than the sweep-\nfrequency one (both modes are of course less sensitive\nthan standard FMR technique where resonant microwave\ncavities are used). However, contrary to the sweep-field\nmode, the sweep-frequency mode presents the advantage\nof performing spectra without disturbing the magneti-\nzation distribution of the studied systems [22], provided\nthat the rf field ~hrfamplitude is weak enough (which is\nthe case here). These two modes have been used in the\npresent study to characterize the indirect magnetoelec-\ntric field and to probe the voltage induced magnetization\nrotation.\nA beforehand characterization of the magnetic prop-\nerties of the heterostructure has been performed in the\nbenchmark state (i. e. absence of applied voltage) in\norder to measure the effective magnetization ( Meff) and\nthe gyromagnetic factor ( \r) as well as the unavoidable\nuniaxial anisotropy ( Hu) due to in-plane non-equibiaxial\nresidual stresses. This initial anisotropy is attributed to\nimperfect flatness of the Kapton substrate during deposi-arXiv:1507.03308v1 [cond-mat.mtrl-sci] 13 Jul 20152\nxM\nHM\nH V Piezoelectric actuator(1cm)Kapton® substrate (125µm)Co2FeAl(25nm)Ta(5nm)\na) b)\n400 800 1200 1600 20004812\n0 90 180 270 3601.141.161.18\nFrequency (GHz)\nH=90\nH=0\nAppliedfield(Oe)Hres(kOe)H(degree)y\nc) d)\nFigure 1: a) Depiction of the cross section view of the in-\nvestigated system illustrating the CFA/Kapton ®system de-\nposited onto a piezoelectric actuator. b) Top view sketch of\nthe system showing the different angles used in the text. c)\nAngular dependence of the resonance field measured at 10\nGHz. d) Experimental and calculated frequency dependence\nof the uniform mode measured along the easy and hard axis\n('H= 90and'H= 0).\n01002003004000 50 100 150 200\n1000 1200 1400 1600‐100‐50050100Appliedvoltage(V)\nExperiments\nLinearfitHR(Oe)\nAppliedfield(Oe)Intensity(µV)\n200V 100V 0Vx\nHH V\ny°=90H °=90H\nFigure 2: Up-graph: Resonance field shift ( \u000eHres =\nHres(V)\u0000Hres(0)) variation as function of the applied volt-\nage. Circles refer to experimental data while solid line repre-\nsents to the best fit. Down-graph: typical MS-FMR spectra\nrecorded at different applied voltage (0 V, 100 V and 200 V).\ntion and/or cementation on the actuator [9]. The quan-\ntitative evaluation of this initial uniaxial anisotropy is\nperformed through the angular dependence of the uni-\nform mode resonance field (see Figure 1-c)) measured at\nf0= 10GHz [15]. Furthermore, one can note that this\nanisotropy is aligned along the y-axis since the resonance\nfield minimum reaches a minimum at 'H= 90°. In ad-\ndition, the frequency dependencies of the uniform mode,\nmeasured along the easy ( 'H= 90°) and hard ( 'H= 0°)\nxyrfh\nMoutP\ninPH\nxyrfh\nM\nMoutPoutP\ninPinPHH\nx\ny\ninPoutP\nMrfh\nH x\ny\ninPinPoutPoutP\nMMrfh\nHHa) b)\nc)\n6.4 6.6 6.8 7.0 7.2‐10‐50510\nHighsensitivity\nLowsensitivity\nLorentzian derivative fits\n6.4 6.6 6.8 7.0 7.2‐0.50.00.5\nFrequency (GHz)Differentiial FMRabsorption (V)Figure 3: a) Sweep-frequency FMR spectra obtained accord-\ning to the two configurations depicted in b) and c) where Pin\nandPoutaretheinjectedandtransmittedradiofrequencycur-\nrent. The only difference comes from the pumping field ( ~hrf)\nwhich either parallel (red squares) or perpendicular (blue cir-\ncles) to~Mdirection giving rise to the high and low sensitivi-\nties geometries. The inset shows a zoom of the marked area.\nNote that a saturating field of 500 Oe was applied to ensure a\nuniform magnetization distribution. The solid lines show fits\nto the first derivative of the standard Lorentzian curve.\naxes are shown in Figure 1-d). The full lines in Figure\n1-c) and -d) are the best fits to the experimental data\nand allowed the determination of the following parame-\nters:Meff= 720emu.cm\u00003,\r= 1:835\u0002107Hz.Oe\u00001\nandHu\u001830Oe, which are close to the values previously\nobtained for similar films [16].\nSweep-field MS-FMR spectra have been recorded as a\nfunction of the applied voltage (for f0= 10GHz). This\nlatter was varied from 0 V to 200 V by 10 V steps. Figure\n2 presents typical spectra measured at different applied\nvoltages (0V, 100 V and 200 V) with an applied mag-\nnetic field along yaxis ('H= 90°, i. e. along the ini-\ntial easy axis). A shift of the resonance field, defined as\n\u000eHres=Hres(V)\u0000Hres(0), isclearlyobserved. Thevolt-\nage dependence of \u000eHresis shown in Figure 2, where the\nsolid line corresponds to a linear fit showing that \u000eHresis\npositive and almost proportional to V. This means that\ntheydirection gradually becomes a hard magnetization\naxis, or a less easy one for the magnetization direction\n[11, 15, 17]. This is due to the positive magnetostriction\ncoefficient of Co 2FeAl[16, 17] and to the nearly uniaxial\ntensile stress (i. e. \u001bxx(V)>0) along the xdirection\napplied by the piezoelectric actuator to the film (for this\nrange of applied voltage: 0 to 200V[23]). Note that a non\nlinear and hysteretic behavior is observed if the voltage\nis applied backward from 200 V to 0 V (not shown here)\ndue to the intrinsic properties of the piezoelectric mate-\nrial [9–11, 15, 23]. It should be noted that the \u000eHres(V)\ncurve presented in Figure 2 has been measured several\ntimes and all the obtained curves were superimposed.\nFurthermore, the resonance field shift can also be viewed\nas an indirect magnetoelectric field Hme(V) =\u000eHres(V)\naligned along xdirection, i. e. perpendicular to the ini-3\ntial easy axis induced by the deposition conditions. The\nevaluation of the voltage dependence of Hmewill serve\nto quantitatively study the magnetization rotation using\nthe sweep-frequency mode of the MS-FMR setup.\nThe sweep-frequency-mode has been used to probe the\nmagnetization direction as function of the applied volt-\nage by measuring both the resonance frequency and the\nspectra amplitudes. For this purpose, the selective rule\nillustrated in Figure 3 has been be used. In this Fig-\nure, two spectra performed with a similar applied mag-\nnetic field (of 500 Oe) along yare presented; the only\ndifference comes from the direction of the weak rf field\n~hrf(also called pumping field) emanating from the mi-\ncrostripline, which is either parallel ( ~hrfk~M) or perpen-\ndicular (~hrf?~M) to the static magnetization. Indeed,\nthe efficiency of the pumping field is maximum when\n~hrf?~M(Figure 3-b) and minimum when ~hrfk~M(Fig-\nure 3-b). In the present experimental conditions, a factor\n20 is found between these two geometries. Furthermore,\nCounil et al.[24, 25] demonstrated both theoretically\nand experimentally that for a pure Lorentzian profile of\nthe signal, the amplitude of this signal is directly propor-\ntional to sin2', where'is the angle between the pump-\ning field and the static magnetization direction. There-\nfore, the voltage-induced rotation of the magnetization\nhas been probed by measuring the evolution of the rel-\native amplitude of the signal as function of the applied\nvoltage.\nFor this purpose, the pumping field was applied paral-\nlel toxdirection (~hrfk~ x), i. e. perpendicular to the\ninitial easy axis (see Figure 1). In order to ensure an ini-\ntial uniform magnetization along the yaxis (initial easy\naxis), a small bias field of 40 Oe was applied. In this\ncondition and in absence of applied voltage, ~Mis sup-\nposed to be uniform and aligned along yaxis. Then, the\nvoltage applied to the actuator, in steps of 10 V, leads\nto the development of an indirect magnetoelectric filed\nparallel to the xaxis, which tends to align ~Malong this\naxis. The normalized component of the magnetization\nalong the indirect magnetoelectric field mx(Hme)is rep-\nresented in Figure using the following formula deduced\nfrom the model presented by Counil et al.[24]:\nmx(Hme) =cos\u0010\narcsin(p\nAN(V)\u0011\ncos\u0010\narcsin(p\nAN(150V)\u0011(1)\nwhereAN(150V)andAN(V)are the normalized (by\nthe0Vamplitudespectrum)amplitudeofthe150Vspec-\ntrum and of the applied voltage spectra, respectively.\nThe red solid line in Figure corresponds to the theo-\nretical value of the normalized x-component of the mag-\nnetization as function of V. One can note that the rota-\ntion is complete at around 40 V ( Hme= 70Oe) where\nthe magnetoelectric field totally compensates the small\nbias field (Hb= 40Oe) and the initial anisotropy field\n(30 Oe). The confrontation between the model and the\n0 25 50 75 100 125 15001234\n0.000.250.500.751.000 40 80 120 160 200 240Frequency (GHz)\nAppliedvoltage(V)\nM(Hme)/MSHme(Oe)\ny\nx\ninPoutP\nM\n0V\nrfhy\nxy\nx\ninPinPoutPoutP\nM\n0VMM\n0V\nrfh\nrfh\ninPoutP\nM\n20V\nrfh\ninPinPoutPoutP\nMM\n20V\nrfh\nrfh\ninPoutP\nM\nrfh 40V\ninPinPoutPoutP\nM\nM\nrfh\nrfh 40VFigure 4: Voltage dependencies of the uniform mode fre-\nquency (blue symbols) and of the normalized x-component\nof the magnetization mx(Hme)with a pumping field applied\nalongx. Themx(Hme)values have been deduced from equa-\ntion1. Thesolidlinesarecalculatedusingequation2withthe\nfollowing parameters: Meff= 720emu.cm\u00003,\r= 1:835\u0002107\ns\u00001,Hu= 30Oe,Hb= 40Oe. Note that the Hmeis deduced\nfrom Figure 2. The sketches show how the magnetization\ndirection rotates as function of the applied voltage.\nexperimental data proves that a voltage induced mag-\nnetization rotation has been performed. Another proof\nof the effective rotation of the magnetization comes from\nthe voltage dependence of the uniform mode frequency as\nfunctionoftheappliedvoltage. ThebluesolidlineinFig-\nure refers to the simulation of this frequency being given\ntheHme(V)dependence previously determined (see Fig-\nure 2). This dependence highlights the well-known soft-\nening of the frequency when the magnetization rotates\nfrom a hard axis to an easy one [26]; the frequency mini-\nmum directly gives the voltage at which the system is in-\nplane isotropic (i. e. where Hmecompensates the initial\nanisotropy field and the bias field). Both the experimen-\ntal frequency and the amplitude dependencies are well\nfitted by the solid lines and they clearly show a 90 degree\nof the magnetization direction. The theoretical values\nhave been calculated by using a magnetic energy den-\nsity of the film where the indirect magnetoelectric field is\nmodeledbyavoltagedependentuniaxialanisotropyfield.\nIn these conditions, the following analytical expression of\nthe frequency can be derived from the Smit-Beljers equa-\ntion [15]:\n2\u0019f\n\r=p\n(Hme(V)\u0000Hu) cos 2'+Hbsin'\u0002\nq\n4\u0019Ms+Hbsin'+Husin2'+Hme(V) cos2'\n(2)4\nwhere'is evaluated at each increment of the applied\nvoltage. Finally, it should be noted that the spectra pro-\nfiles were not always close to a Lorentzian one, especially\nat high applied voltages, which make this modeling really\nrobust [24, 25].\nIn summary, the indirect magnetoelectric coupling in\nartificial heterostructure was used to rotate the mag-\nnetization direction. MS-FMR was used in two differ-ent modes to: i) quantitatively measure the voltage-\ndependence of the indirect magnetoelectric field (which\nis strain-induced by the piezoelectric actuator) and to ii)\nobserve the rotation of the magnetization. The voltage\ndependence of the Hmewas found to be almost linear (in\nthe range 0V-200V) and was used during the frequency\nand amplitude variations of the sweep-frequency spectra\nmodeling.\n[1] N. A. Spaldin and M. Fiebig, Science, 309, 391 (2005)\n[2] S. -W. Cheong and M. Mostovoy, Nature Materials 6, 13\n(2007)\n[3] R. Ramesh and N. A. Spaldin, Nature Materials 6, 21\n(2007)\n[4] M. Bibes and A. Barthélémy, Nature Materials 7, 425\n(2008)\n[5] W. Eerenstein, N. Mathur, and J. Scott, nature 442, 759\n(2006)\n[6] Ce-Wen Nan, M. I. Bichurin, S. Dong, D. Viehland, and\nG. Srinivasan, Journal of Apllied Physics 103, 031101\n(2008)\n[7] N. X. Sun and G. Srinivasan, Spin, 2, 1240004 (2012)\n[8] J. Ma, J. Hu, Z. Li, Ce-Wen Nan, Advanced Materials\n23, 1062 (2011)\n[9] M. Gueye, F. Zighem, D. Faurie, M. Belmeguenai and S.\nMercone, Applied Physics Letters 105, 052411(2014)\n[10] M. Weiler, A. , S. Geprägs, M. Althammer, M. Opel, C.\nBihler, H. Huebl, M. Brandt, R. Gross, and S. Goennen-\nwein, New Journal of Physics 11, 013021 (2009)\n[11] F. Zighem, D. Faurie, S. Mercone, M. Belmeguenai, H.\nHaddadi, Journal of Applied Physics 114, 073902 (2013)\n[12] J. H. Park, J.-H. Lee, M. G. Kim, Y. K. Jeong, M.-A.\nOak, H. M. Jang, H. J. Choi, and J. F. Scott, Physical\nReview B 81, 134401 (2010)\n[13] C. Thiele, K. Dörr, O. Bilani, J. Rödel, and L. Schultz,\nPhysical Review B 75, 054408 (2007)\n[14] C. Bihler, M. Althammer, A. Brandlmaier, S. Geprägs,\nM.Weiler, M. Opel, W. Schoch, W. Limmer, R. Gross,\nM. S. Brandt, and S. T. B. Goennenwein, Phys. Rev. B\n78, 045203 (2008).\n[15] F. Zighem, A. El Bahoui, J. Moulin, D. Faurie, M.\nBelmeguenai, S. Mercone and H. Haddadi, Journal ofApplied Physics 116, 123903 (2014)\n[16] M. Gueye, B. M. Wague, F. Zighem, M. Belmeguenai,\nM. S. Gabor, T. Petrisor, C. Tiusan, S. Mercone, and D.\nFaurie, Applied Physics Letters 105, 062409 (2014).\n[17] S. Li, J. Xu, Q. Xue, H. Du, Q. Li, C. Chen, R. Yang, S.\nXie, M. Liu, T. Nan, N. X. Sun, and W. Shao, Journal\nof Applied Physics 117, 17B722 (2015)\n[18] M. Belmeguenai, H. Tuzcuoglu, M. Gabor, T. Petrisor\nJr, C. Tiusan, D. Berling, F. Zighem, T. Chauveau, S.\nChérif, and P. Moch, Physical Review B 87, 184431\n(2013)\n[19] W. Wang, H. Sukegawa, R. Shan, S. Mitani, and K. In-\nomata, Applied Physics Letters 95, 182502 (2009)\n[20] M. S. Gabor, M. Belmeguenai, F. Zighem, S. M. Chérif,\nT. Petrisor Jr, T. Petrisor, C Tiusan and M Hehn, Spin,\n4, 1440022 (2014)\n[21] F.Zighem,Y.Roussigné,S.M.ChérifandP.Moch,Jour-\nnal of Physics: Condensed Matter 20, 125201 (2008)\n[22] J. Ben Youssef, N. Vukadinovic, D. Billet and M.\nLabrune, Physical Review B 69, 174402 (2004)\n[23] F. Zighem, M. Belmeguenai, D. Faurie, H. Haddadi and\nJ. Moulin, Review of Scientific Instruments, 85, 103905\n(2014)\n[24] G. Counil, Joo-Von Kim, T. Devolder, C. Chappert, K.\nShigeto and Y. Otani, Journal of Applied Physics 95,\n5646 (2004)\n[25] G. Counil, J.-V. Kim, T. Devolder, P. Crozat, C. Chap-\npert and A. Cebollada, Journal of Applied Physics 98,\n023901 (2005)\n[26] F.Zighem,Y.Roussigné,S.M.ChérifandP.Moch,Jour-\nnal of Physics: Condensed Matter 19, 176220 (2007)" }, { "title": "1509.01807v1.Study_of_spin_dynamics_and_damping_on_the_magnetic_nanowire_arrays_with_various_nanowire_widths.pdf", "content": "1 \n Study of spin dynamics and damping on the magnetic nanowire arrays \nwith various nanowire widths \n \nJaehun Cho a, Yuya Fujii b, Katsunori Konioshi b, Jungbum Yoon c, Nam -Hui Kim a, Jinyong Jung a, \nShinji Miwa b, Myung -Hwa Jung d, Yoshishige Suzuki b, and Chun -Yeol You a,* \n \na Department of Physics, Inha University , Inch eon, 402-751, South Korea \nb Graduate School of Engineering Science, \nOsaka University, Toyonaka, Osaka 560 -8531, Japan \nc Department of Electrical and Computer Engineering , \nNational University of Singapore , Singapore 117576 \nd Department of Physics, Sogang University, Seoul, 121 -742, South Korea \n \nAbstract \nWe investigate the spin dynamics including Gilbert damping in the ferromagnetic nanowire \narray s. We have measured the ferromagnetic resonance of ferromagnetic nanowire arrays \nusing vector -network analyzer ferromagnetic resonance (VNA -FMR) and analyzed the results \nwith the micromagnetic si mulations . We find excellent agree ment between the experimental \nVNA -FMR spectra and micromagnetic simulations result for various applied magnetic fields . \nWe find that the demagnetization factor for longitudinal conditions, Nz (Ny) increases \n(decreas es) as decreasing the nanowire width in the micromagnetic simulation s. For the \ntransverse magnetic field , Nz (Ny) increas es (decreas es) as increasing the nanowire width . We \nalso find that t he Gilbert damping constant increases from 0.018 to 0.051 as the incr easing \nnanowire width for the transverse case , while it is almost constant as 0.021 for the \nlongitudinal case . \n 2 \n * Corresponding author. FAX: +82 32 872 7562. \nE-mail address: cyyou@inha.ac.kr \nKeywords : Nanowires , Ferromagnetic Resonance , Micromagnetic Simulations , Gilbert \ndamping \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n 3 \n \nFerromagnetic nanostructures have recently attracted much interest for the wide potential \napplications in high density spintronic information storage , logic devices and various spin \norbit torque phenomena .1,2,3,4,5 It is well known that the detail spin dynamics of nanostructure \nis far from the one of the bulk’s because of many reasons, different boundary conditions, \nchanges of the magnetic properties including the saturat ion magnetization, anisotropy energy, \nand exchange stiffness constant, etc. Since the magnetic properties are usually sensitive \nfunctions of the sample fabrication conditions, it has been widely accepted that the detail \nsample fabrications are also importa nt in the study of spin dynamics. However, the relatively \nless caution has been made for the boundary conditions of the spin dynamics in the \nnanostructure. \nIn the spin transfer torque magnetic random access memory (STT -MRAM), the magnetic \ndamping constant is important because the switching current density is proportional to the \ndamping constant .6 In the nanowire, damping constant also plays crucial role in the spin \ndynamics including domain wall motion with magnetic field7 and spin transfer torque .8 \nFurthermore, it is the most important material parameter in spin wave (SW) dynamics .9 \nDespite of the importance of the damping constant, many studies about spin dynamics in \nferromagnetic nanowires have not taken into account the damping constant properly .10,11,12 \nOnly a few studies paid attention to the magnetic damping in the nanowires spin 4 \n dynamics .13,14 \nIn this study, arrays of CoFeB nanowire s are prepared by e -beam lithography , and they are \ncovered coplanar wave -guide for the ferromagnetic resonance (FMR) measurement as shown \nin Fig. 1 . We measured FMR signal with longitudinal (wire direction) and transverse \nmagnetic field s in order to investigate the spin dynamics with different boundary conditions. \nAlso w e extract Gilbert damping constant using micromagnetic simulations with the different \napplied magnetic field directions in various nanowire arrays . We find the damping constant \ndecreas es with increasing the nanowire width for the transverse magnetic field with constant \ninput damping consta nt in micromagnetic simulations, while we obtain almost constant \ndamping constant for the longitudinal field. \nThe film s were prepared using DC magnetron sputtering. The stack s consist of Ta (5 \nnm)/Co 16Fe64B20 (30 nm)/Ta (5 nm) on single crystal MgO (001) substrate s. The film s are \npatterned as 100 -nm-width wire array s with 200 -nm-space each wires using e -beam \nlithography and an Ar ion milling technique as shown in Fig. 1. The width is determined with \na scanning electron microscope (SEM). These nanowire arrays are covered by coplanar wave \nguide in order to characterized with the Vector Network Analyzer ( VNA )-FMR technique \ndescribed elsewhere .15 We prepare nanowire arrays as shown in Fig. 1 , and external DC \nmagnetic field direction for FMR measurement is also depicted. \nWe use VNA -FMR spectra to measure imaginary parts of the susceptibility of the samples.16 5 \n The measured imaginary parts of the susceptibility raw data are calibrated with the careful \ncalibration procedures .16 The calibrated imaginary parts of the susceptibility are shown in Fig \n2(a) and (b) for an applied magnetic field at 0.194 T for the nanowire arrays . The un -\npatterned thin film is also examined for the reference. We find two resonance frequencies, \n17.2 and 26 .4 GHz, as shown in Fig. 2(a) for the nanowire array, while there is only one peak \nat 16.8 GHz for the un-patterned thin film as shown in Fig 2(b). We believe that the smaller \npeak (17.2 GHz) in Fig. 2(a) is originated from the un-patterned part of the nanowire array, \nbecause the frequency is closed to the un -patterned thin film’s peak (16.8 GHz). Probably, the \nun-patterned part of the nanowire is formed due to poor e -beam lithography processes. On the \nother hand, t he resonance f requency near 26.0 GHz is calculated from micromagnetic \nsimulation at an applied magnetic field at 0.200 T , as shown in Fig. 2 (c). We clarif y the \nsource of the main peak (26.4 GHz) is nanowire arrays by using micromagnetic simulation. \nThese two peaks name d as the uniform FMR mode (smaller peak position) and nanowire \nmode (higher peak position). \nIn order to determine the saturation magnetization, the resonance frequencies are measured \nas a function of the applied magnetic field, and the results are fitted with the Kittel ’s \nequation .17 This equation employs the corresponding demagnetization factors of Nx = 0, Ny = \n0 and Nz = 1 for un -patterned film, when applied magnetic field H is x- direction with \nfollowing equations , 6 \n \n 2y x s z x s f H N N M H N N M\n \n. (1) \n \n Here, is the gyromagnetic ratio, H is the applied magnetic field, Ms is saturated \nmagnetization, Nx, Ny, and Nz are the demagnetization factor s applying the cyclic permutation \nfor the applied magnetic field direction . \nThe micromagnetic simulations are performed by using the Objective -Oriented -\nMicroMagnetic Framework (OOMMF)18 with 2-dimensional periodic boundary condition \n(PBC ).19 We select a square slat of 100 nm × 100 nm × 30 nm nanowire separated 200 nm in \ny- direction with a cell size of 5 nm × 5 nm × 30 nm. The material parameters of CoFeB used \nin our simulation are summarized as follows: Ms = 15.79 × 105 A/m, the exchange stiffness \n1.5 × 1011 J/m, the gyromagnetic ratio 2.32 × 1011 m/(A ∙s) and we ignore the magneto -\ncrystalline anisotropy. In this simulation, the Gilbert damping constant of 0.0 27 is fixed. The \nsaturation magnetization and Gilbert damping constant are determined by using VNA -FMR \nmeasurement for un -patterned thin film . For t he exchange stiffness constant, experimentally \ndetermined values are range of 0.98 to 2.84 × 1011 J/m which value has dependence on the \nfabrication processes20 and composition of ferromagnetic materials ,21 while we have picked \n1.5 × 1011 J/m as the exchange stiffness constant . The determination method of Gilbert \ndamping constant will be described later. 7 \n In order to mimic FMR experiments in the micromagnetic simulations , a “sinc” function\n0 0 0 ( ) sin 2 / 2y H HH t H f t t f t t \n, with H0 = 10 mT, and field frequency fH = 45 \nGHz, is applied the whole nanowire area.22 We obtain the FMR spectra in the corresponding \nfrequency range from 0 to 45 GHz . The FMR spectra due to the RF -magnetic field are \nobtained by the fast Fourier transform (FFT) of stored My(x) (x, y, t ) in longitudinal (transverse) \nH0 field. More details can be described elsewhere .23 \nThe closed blue circles in Fig. 3 is the calculated values with the fitting parameter using Eq. \n(1) which are fitted with the experimental data of un -patterned thin film. The obt ained Ms is \n15.79 × 105 A/m while gyromagnetic ratio is fixed as 2.32 × 1011 m/(A∙s) . The obtained Ms \nvalue is similar with vibrating sample magnetometer method24 which CoFeB structure has Ta \nbuffer layer. The resonance frequencies of uniform FMR mode in nanowire arrays are plotted \nas open red circles in Fig. 3. The resonance frequencies of uniform FMR mode is measured \nby VNA -FMR are agreed well with resonance freq uency of un-patterned thin film measured \nby VNA -FMR. In Fig. 3, the applied field dependences of the resonance frequencies \nMeasured by VNA -FAM for the nanowire are plotted as open black rectangular , along with \nthe result of micromagnetics calculated with E q. (1) as depicted closed black rectangular . It is \nalso well agreed with the experimental result in nanowire mode and micromagnetic \nsimulation result in the nanowire arrays. \nIn order to reveal the effect s of spin dynamics properties with various nanowire width s, we 8 \n perform micromagnetic simulat ions. The nanowire width s are varied from 50 to 150 nm in \n25-nm step for fixed 200 -nm-space with PBC , it causes changes of the demagnetization \nfactor of the nanowire . In Fig. 4 (a) shows the nanowire width dependences of the resonance \nfrequencies for the longitudinal magnetic field (open symbols) along with the resonance \nfrequencies calculated with Eq. (1) (solid lines) . The demagnetization factors can be \ndetermined by fitting Eq. (1) while Nx is fixed as 0 to represent infinitely long wire . The \nagreements between the results of micromagnetic simulations (open circles) and Eq. (1) \n(solid lines) are excellent. \nFor the transverse magnetic field, the direction of applied magnetic field is y - axis, Eq. (1) \ncan be rewritten as follows: \n \n 2x y s z y s f H N N M H N N M\n \n. (2) \n \nIn this equation, we use the relation of demagnetization factors , \n1x y zN N N , in \norder to remove uncertainty in the fitting procedure . In the transverse field, the \ndemagnetization factors are determined by Eq. (2). The resonance frequencies for transverse \nmagnetic field which are obtained by micromagnetic simulation (open circles) and \ncalculated by Eq. (2) (solid lines) as a function of the appl ied magnetic field with various \nnanowire width are displayed in Fig. 4(b). The longitudinal case, when the field direction is 9 \n easy axis, they are saturated with small field. However, the transverse case, when the field \ndirection is hard axis, certain amoun t of field is necessary to saturate along the transverse \ndirections. The narrower nanowire, the larger field is required as shown in Fig. 4 (b). \nFig. 5(a) and (b) show the changes of demagnetization factors in longitudinal and transverse \nmagnetic fields as a functi on of the nanowire width , respectively. The demagnetization \nfactors play important role in the domain wall dynamics, for example the Walker breakdown \nis determined by the demagnetization factors .25 Furthermore, they are essential physical \nquantities to analyze the details of the spin dynamics. It is clear ly shown that the Nz (Ny) \nincrease s (decrease s) with increasing the nanowire width in longitudinal magnetic field. For \nthe transverse magnetic field, Nz (Ny) increase s (decrease s) with increasing the nanowire \nwidth , during Nx is almost zero value. The demagnetization factors both longitudinal and \ntransverse have similar tendency with the effective demagnetization factors of dynamic \norigin26 and the static demagnetization factors for the prism geometry.27 \nNow, let us discuss about the Gilbert damping constant . The relation of the full width and \nhalf maxim a (f) of a resonance peak s as a function of applied field are shown in Fig. 6 for \nlongitudinal (a) and transverse (b). The f is given by15: \n \n,\n,2\n22xy\ns z ex yxN\nf H M N N f\n \n. (3) \n 10 \n where, fex is the extrinsic line width contributions , when the applied magnetic field is x-(y-\n)axis for longitudinal (transverse) case . The symbol s are the result s of the micromagnetic \nsimulations and the solid lines are the fitting result of Eq. (3) . We use pre -determined \ndemagnetization factors (Fig. 5) during fitting procedures, and the agree ments are excellent. \nWe have plotted the Gilbert damping constant as a function of the wavevector in nanowire \nwidth (\n/ qa , a is the nanowire width ) in Fig. 7. The black open rectangles are data \nextracted from the transverse field and the red open circles are longitudinal field data. We \nfind that the Gilbert damping constant varied from 0.051 to 0.018 by changing the \nwavevector in nanowire width in transverse field. On the other hand, lon gitudinal field case \nthe damping constant is almost constant as 0.021. Let us discuss about the un -expected \nbehavior of the damping constant of transverse case. T he wire width acts as a kind of cut -off \nwavelength of the SW excitations in the confined geome try. SWs whose wavelength are \nlarger than 2 a are not allowed in the nanowire. Therefore, only limited SW can be excited for \nthe narrower wire, while more SW can be existed in the wider wire. For example, we show \ntransverse standing SW as profiled in the inset of Fig. 6 for 150 -nm width nanowire in our \nmicromagnetic simulations. More possible SW excitations imply more energy dissipation \npaths, it causes larger damping constant. For narrower nanowire (50 -nm), only limited SWs \ncan be excited, so that the damping constant is smaller. However, for the limit case of infinite \na case, it is the same with un -patterned thin films, there is no boundar y so that only uniform 11 \n mode can be excited, the obtained damping constant must be the input value. \nIn summary, the VNA -FMR experiments is employed to investigate the magnetic properties \nof CoFeB nanowire arrays and the micromagnetic simulations is proposed to understand the \nmagnetic properties including Gilbert damping constant of various CoFeB nanowire arrays \nwidth. We f ind that the demagnetization factors are similar with the dynamic origin and static \nfor the prism geometry. The wire width or SW wavevector dependent damping constants can \nbe explained with number of SW excitation modes. \n \nACKNOWLEDGMENTS \nThis work was supported by the National Research Foundation of Korea (NRF) Grants (Nos. \n616-2011 -C00017 and 2013R1A12011936 ). \nReferences \n \n1 S. S. P. Parkin, M. Hayashi, and L. Thomas, Science 320, 190 (2008) . \n2 D. A. Allwood, G. Xiong, C. C. Faulkner, D. Atkinson, D. Petit, and R. P. Cowburn, \nScience 309, 1688 (2005) . \n3 I. M. Miron, G. Gaudin, S. Aufftet, B. Rodmacq, A. Schuhl, S. Pizzini, J. V ogel and P. \nGambardella, Nature Materials 9, 230 (2010) . \n4 H-R Lee , K. Lee, J. Cho, Y . -H. Choi, C. -Y . You, M. -H. Jung, F. Bonell, Y. Shiota, S. Miwa \nand Y . Suzuki, Sci. Rep. 4, 6548 (2014). \n5 J. Cho, et al. Nat. Commun. 6, 7635 (2015). \n6 S. Ikeda , K. Miura, H. Yamamoto, K. Mizunuma, H. D. Gan, M. Endo, S. Kanai, J. \nHayakawa, F. Matsukura and H. Ohno , Nat. Mater. 9, 721 (2010). \n7 J.-S. Kim et al. , Nat. Commun . 5, 3429 (2014). \n8 L. Thomas, R. Moriya, C. Rettner and S. S. P. Parkin , Science 330, 1810 (2010). 12 \n \n9 J.-S. Kim, M. Stark, M. Klaui, J. Yoon, C. -Y . You, L. Lopez -Diaz and E. Martinez , Phys. \nRev. B 85, 174428 (2012). \n10 B. Kuanr, R. E. Camleym, and Z. Celinski, Appl. Phys. Let t. 87, 012502 (2005) . \n11 J. P. Park, P. Eames, D. M. Engebretson, J. Berezovsky, and P. A. Crowell, Phys. Rev. Le tt. \n89, 277201 (2002). \n12 L. Kraus, G. Infante, Z. Frait and M. Vazquez, Phys. Rev. B 83, 174438 (2011) . \n13 C. T. Boone , J. A. Katine, J. R. Childress, V . Tiberkevich, A. Slavin, J. Zhu, X. Cheng and \nI. N. Krivorotov , Phys. Rev. Let t. 103, 167 601 (2009) . \n14 M. Sharma, B. K. Kuanr, M. Sharma and Ananjan Nasu, IEEE Trans. Magn 50, 4300110 \n(2014). \n15 D.-H. Kim, H .-H. Kim, and Chun -Yeol You, Appl. Phys. Lett. 99, 072502 (2011). \n16 D.-H. Kim, H. -H. Kim, Chun -Yeol You and Hyungsuk Kim, J. Magn etics 16, 206 (2011). \n17 C. Kittel, Introduction to Solid State Physics, 7th ed., pp. 504, (1996) . \n18 M. J. Donahue and D. G. Porter: OOMMF User’ s Guide : Ver. 1.0, NISTIR 6376 (National \nInstitute of Standards and Technology, Gaithersburg, Maryland, United States, 1999 ). \n19 W. Wang, C. Mu, B. Zhang, Q. Liu, J. Wang, D. Xue, Comput. Mater. Sci. 49, 84 (2010) . \nSee: http://oommf -2dpbc.sourceforge.net. \n20 J. Cho, J . Jung, K .-E. Kim, S .-I. Kim, S .-Y. Park, M .-H. Jung, C .-Y. You, J. Magn. Magn. \nMater. 339, 36 (2013). \n21 C. Bilzer, T. Devolder, J -V . Kim, G. Counil, C. Chappert, S. Cardoso and P. P. Feitas , J. \nAppl. Phys. 100, 053903 (2006). \n22 K.-S. Lee, D. -S. Han , S.-K. Kim, Phys. Rev. Let t. 102, 127202 (2009). \n23 J. Yoon, C. -Y . You, Y . Jo, S. -Y. Park, M. H. Jung , J. Korean Phys. Soc. 57, 1594 (2010) . \n24 Y . Shiota, F. Bonell, S. Miwa, N. Muzuochi, T. Shinjo and Y . Suzuki , Appl. Phys. Le tt. 103. \n082410 (2013), \n25 I. Purnama, I. S. Kerk, G. J. Lim and W. S. Lew , Sci. Rep. 5, 8754 (2015). \n26 J. Ding, M. Kostylev, and A. O. Adeyeye, Phys. Rev. B. 84, 054425 (2011) . \n27 A. Aharoni, J. Appl. Phys. 83, 3432 (2011) . 13 \n Figure Captions \n \nFig. 1. Measurement geometry with SEM image s of the 100 -nm-width nanowires with a gap \nof 200 nm between nanowires . The longitudinal nanowire arrays are shown. After the \nnanowire patterns have been defined by e -beam lithography, they are covered by co -planar \nwave guides. \n \nFig. 2. (a) The measured FMR spectrum of the CoFeB nanowire with H =0.194 T. The red (lower \npeak) and blue (higher peak) arrows indicate t he resonance frequencies of the uniform FMR mode and \nthe nanowire mode, respectively. (b) The measured FMR spectrum of the CoFeB thin film with H \n=0.194 T. (c) Simulated FMR spectrum of the CoFeB nanowire with H= 0.200 T. \n \nFig. 3. Measured and calculated FMR frequencies with the applied magnetic field for 100 -\nnm-width nanowire. The open black rectangles are nanowire mode and open red circles are \nthe uniform FMR mode for CoFeB thin film. The closed black rectangles are calculated by \nOOMMF and the closed blue circles are theoretical ly calculated by Eq. (1) using fitted \nparameters form un -patterned film . \n \nFig. 4. Variation of resonanc e frequencies with the applied magnetic field for the different \nPBC wire width for (a) longitudinal field and (b) transverse field. Inset: The geometry of 2 -\ndimensional PBC micromagnetic simulation with nanowire width a and a gap of 200 nm \nbetween nanowire s. The black open rectangles, red open circles, green open upper triangles, \nblue open down triangles, cyan open diamonds represent as nanowire width as 50 nm, 75nm, \n100 nm, 125nm, and 150 nm, respectively. \n \nFig. 5. Demagnetization factor with PBC wire widt h for (a) longitudinal and (b) transverse \nfield. The black open circles, red open rectangles, blue open upper triangles represent as \ndemagnetization factors, Ny, Nz, and Nx, respectively. \n \nFig. 6. Full width and half maxim a with the applied magnetic field for (a) longitudinal and (b) \ntransverse field. The black open rectangles, red open circles, green open upper triangles, blue \nopen down triangles, cyan open diamonds represent as nanowire width as 50 nm, 75nm, 100 \nnm, 125nm, and 150 nm, respectively. \n \nFig. 7. Damping constants with wavevector for transverse ( the black open rectangles ) and \nlongitudinal ( the red open circles) field with errors. The black line is the input value which is \ndetermined from un -patterned film. Inset presents the profile of the trans verse spin density as SWs. \n Fig. 1 \n \n \n \n \n Fig. 2. \n \n \n \nFig. 3. \n \n \n \n \nFig. 4 \n \n` \n \n \n \nFig. 5 \n \n \n \nFig. 6 \n \n \n \nFig. 7 \n \n \n" }, { "title": "2001.02895v1.Hybridization_between_the_ligand__p__band_and_Fe_3_d__orbitals_in_the_p_type_ferromagnetic_semiconductor__Ga_Fe_Sb.pdf", "content": " 1 Hybridization between the ligand p band and Fe-3d orbitals in the p-type ferromagnetic semiconductor (Ga,Fe)Sb Takahito Takeda1, Masahiro Suzuki2, Le Duc Anh1,3, Nguyen Thanh Tu1,4, Thorsten Schmitt5, Satoshi Yoshida6, Masato Sakano6, Kyoko Ishizaka6, Yukiharu Takeda7, Shin-ichi Fijimori7, Munetoshi Seki1,8, Hitoshi Tabata1,8, Atsushi Fujimori2,9, Vladimir N. Strocov5, Masaaki Tanaka1,8, and Masaki Kobayashi1,8 1Department of Electrical Engineering and Information Systems, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan 2Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan 3Institute of Engineering Innovation. University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan 4Department of Physics, Ho Chi Minh City University of Pedagogy, 280, Au Duong Vuong Street, District 5, Ho Chi Minh City 748242, Vietnam 5Swiss Light Source, Paul Scherrer Institut, CH-5232 Villigen PSI, Switzerland 6Quantum-Phase Electronics Center and Department of Applied Physics, University of Tokyo, Bunkyo, Tokyo 113-8656, Japan 7Synchrotron Radiation Research Unit, Japan Atomic Energy Agency, Sayo-gun, Hyogo 679-5148, Japan 8Center for Spintronics Research Network, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan 9Department of Applied Physics, Waseda University, Okubo, Shinjuku, Tokyo 169-8555, Japan (Date: January 9, 2019) ABSTRACT (Ga,Fe)Sb is a promising ferromagnetic semiconductor for practical spintronic device applications because its Curie temperature (TC) is above room temperature. However, the origin of ferromagnetism with high TC remains to be elucidated. Here, we use soft x-ray angle-resolved photoemission spectroscopy (SX-ARPES) to investigate the valence-band (VB) structure of (Ga0.95,Fe0.05)Sb including the Fe-3d impurity band (IB), to unveil the mechanism of ferromagnetism in (Ga,Fe)Sb. We find that the VB dispersion in (Ga0.95,Fe0.05)Sb observed by SX-ARPES is similar to that of GaSb, indicating that the doped Fe atoms hardly affect the band dispersion. The Fe-3d resonant ARPES spectra 2 demonstrate that the Fe-3d IB crosses the Fermi level (EF) and hybridizes with the VB of GaSb. These observations indicate that the VB structure of (Ga0.95,Fe0.05)Sb is consistent with that of the IB model which is based on double-exchange interaction between the localized 3d electrons of the magnetic impurities. The results indicate that the ferromagnetism in (Ga,Fe)Sb is formed by the hybridization of the Fe-3d IB with the ligand p band of GaSb. I. INTRODUCTION Ferromagnetic semiconductors (FMSs) are alloy semiconductors in which cation sites are partially replaced by a sizable amount of magnetic impurities, leading to ferromagnetic properties. The ferromagnetism of FMSs is considered to originate from the magnetic interaction between the doped magnetic impurities mediated by the spin of the carriers. This nature is called carrier-induced ferromagnetism1. FMSs have attracted much attention as promising materials to apply for semiconductor spintronic devices, because one can control their magnetic properties by changing the carrier concentration. The p-type Mn-doped III-V FMSs, such as (In,Mn)As2,3,4,5 and (Ga,Mn)As6,7,8, have been intensively studied so far as prototypical FMSs showing carrier-induced ferromagnetism. 3 Spintronics devices based on these FMSs have been fabricated9,10. However, these materials are still seriously problematic for practical applications. Firstly, Mn-doped FMSs are only p-type because the doped Mn atoms act as acceptors in III-V semiconductors. Secondly, the reported maximum values of the Curie temperature (TC) of molecular-beam epitaxy (MBE)-grown Mn-doped FMSs are much lower than room temperature. The highest TC of (Ga,Mn)As reported so far is ~ 200 K11 and that of (In,Mn)As is ~ 90 K12. Recently, Fe-doped III-V FMSs such as n-type (In,Fe)As13, n-type (In,Fe)Sb14, and p-type (Ga,Fe)Sb15 have been successfully grown. Since the doped Fe ions are expected to isovalently substitute for the cation (In3+ or Ga3+) sites of III-V semiconductors as Fe3+, one can independently control the concentrations of Fe ions and, by doping other atoms, carriers in Fe-doped FMSs. Furthermore, the previously reported highest TC values n-type (In0.84,Fe0.16)Sb (335 K)16, (In0.65,Fe0.35)Sb (385 K)17, and p-type (Ga0.8,Fe0.2)Sb (> 400 K)18 are well above room temperature. Considering these advantages, Fe-based FMSs are more promising materials for applications to semiconductor spintronic devices operating at room temperature. Understanding the mechanism of the carrier-induced ferromagnetism in FMSs is 4 important for designing functional FMSs materials and for practical applications. Several theoretical models, such as the Zener’s p-d exchange model and impurity band (IB) model have been proposed so far19. The former, based on a mean-field theory, has been proposed as the itinerant limit where the holes are considered to be nearly free carriers20,21. This model indicates that the Fermi level (EF) is located in the valence band (VB) and that the origin of the ferromagnetism is the p-d exchange interactions between the VB holes and localized 3d electrons of the magnetic impurity. In contrast, the latter model has been proposed as the other limit where the hole carriers are localized around the magnetic impurities22,23. This model indicates that EF is located in the IB and that the ferromagnetism arises from the double-exchange interaction between the localized 3d electrons of the magnetic impurities. As described above, the origin of ferromagnetism in FMSs is considered to be related to its band structure near EF. Experimentally, the electronic states of (Ga,Mn)As in the vicinity of EF have been studied by angle-resolved photoemission spectroscopy (ARPES) to unveil the mechanism of carrier-induced ferromagnetism from the viewpoint of electronic band structure24,25,26,27. Soft x-ray (SX) ARPES measurements have been instrumental to 5 directly access the three-dimensional (3D) band structure and Mn-3d IB of (Ga,Mn)As27. Here, we investigate the electronic structure of the Fe-based FMS (Ga,Fe)Sb to reveal the origin of its high TC, and particularly examine the Fe-3d IB in the vicinity of EF, by SX-ARPES measurements. The SX-ARPES results provide an understanding of the carrier-induced ferromagnetism of (Ga,Fe)Sb. II. EXPERIMENTAL (Ga0.95,Fe0.05)Sb and GaSb thin films a thickness of 30 nm were grown on semi-insulating (SI) GaAs(001) substrates by MBE. To avoid surface contamination, the surfaces of the films were covered with amorphous Sb capping layers. Figure 1(a) shows the schematic sample structure. During the MBE growth, the excellent crystallinity of the samples was confirmed by reflection high-energy electron diffraction. The lattice constant (a) of (Ga0.95,Fe0.05)Sb is 0.608 nm 28. The value of TC of (Ga0.95,Fe0.05)Sb, which was grown by the same condition as the measured one, is about 25 K15. The SX-ARPES experiments were performed at the SX-ARPES end station29 of the ADRESS beamline at the Swiss Light Source. Before the SX-ARPES measurements, the samples were annealed 6 at around 300 oC in the preparation chamber to remove the amorphous Sb capping layer and to expose the clean surface of the samples. The measurements were conducted under an ultrahigh vacuum below 10-10 mbar at a temperature of 12 K, with varying the photon energy (hν) from 500 eV to 1000 eV. The total energy resolution including the thermal broadening was between 50 meV and 200 meV depending on hν. The incident beam with linear-vertical and linear-horizontal polarizations, which respectively correspond to p-polarization and s-polarization configurations30, were used for the measurements. The Fe L2,3 x-ray absorption spectroscopy (XAS) spectra were measured in the total-electron-yield mode. III. RESULTS AND DISCUSSION A. Constant energy mappings and band dispersion around the Γ point Figures 2(a) - 2(c) show out-of-plane (kz-k//) constant-energy mappings for (Ga0.95,Fe0.05)Sb at different binding energy (EB), where the red solid lines represent the Brillouin zone (BZ) as shown Fig. 1(b). Here, kz and k// (= k[-110]) are out-of-plane (Γ-X-Γ) and in-plane (Γ-K-X) momenta, respectively. The observed band dispersion clearly 7 depends on kz and reflects the symmetry of the BZ. Note here that there is no band that is non-dispersive along the kz direction in the mappings, evidencing the absence of surface (two-dimensional) states in the present SX-ARPES data. These observations demonstrate that the SX-ARPES spectra reflect the 3D band dispersion of the (Ga0.95,Fe0.05)Sb thin film. Figures 3(a) and 3(b) show the out-of-plane and in-plane Fermi surface mappings (FSMs) of the (Ga0.95,Fe0.05)Sb thin film, respectively. It should be noted here that the cut taken at hν = 885 eV contains the Γ point, as shown in Fig. 3(a). Then, the Γ-K-X symmetry line is precisely determined from the in-plane k[-110]-k[-1-10] constant-energy mapping taken at hν = 885 eV as shown in Fig. 3(b). Figures 3(c) and 3(d) show EB vs. momentum (k) plots along the Γ-K-X symmetry line with p and s polarizations, respectively. The light-hole (LH) and split-off (SO) bands show up with p polarization in Fig. 3(c), while the heavy-hole (HH) band is clearly visible in the SX-ARPES image taken with s polarization [Fig. 3(d)]. This linear polarization dependence comes from the wave-function symmetry of these bands31. Following the LH-band peaks of energy distribution curves (EDCs) in Fig. 3(e), the top of the LH band 8 is located at 150 meV below EF. Because the band gap (Eg) of (Ga0.95,Fe0.05)Sb is close to that of GaSb (Eg = 812 meV32)33, the observation evidences that (Ga0.95,Fe0.05)Sb is p-type, which is in agreement with the transport measurements15,28,34. To examine the Fe-doping effects on the band dispersion, ARPES measurements on the GaSb thin film has also been conducted as a reference. Figure 4(a) shows EB vs k plots for GaSb along the Γ-K-X symmetry line with p polarizations at the hν of 880 eV. GaSb is p-type because the top of the LH band is close to EF. Comparing the band dispersions between (Ga0.95,Fe0.05)Sb in Fig. 3(c) and GaSb in Fig. 4(a), the band dispersion of (Ga0.95,Fe0.05)Sb is nearly identical to that of GaSb. Figure 4(b) shows the EDCs of GaSb (blue line) and (Ga0.95,Fe0.05)Sb (red line) along the Γ-K-X symmetry line with p polarization. The ARPES spectra of (Ga0.95,Fe0.05)Sb are broader than those of GaSb because of the structural disorder due to the Fe doping and/or because of the Zeeman splitting. However, the peak positions of EDCs of (Ga0.95,Fe0.05)Sb are almost the same as those of GaSb. This result indicates that the band dispersion itself originating from the sp-orbitals of GaSb is hardly affected by doping of Fe atoms. Since the position of EF in 9 (Ga0.95,Fe0.05)Sb is approximately the same as that in GaSb, the Fe atoms would isovalently substitute for the Ga site. B. Fe-3d impurity band Note that the Fe-3d IB is hardly seen in the ARPES spectra of (Ga0.95,Fe0.05)Sb near the Γ point in Figs. 3(c) and 3(d) due to the small amount of Fe atoms, although the band dispersion of the GaSb host has been clearly observed. The energy position of the Fe-3d states is a key to understand the ferromagnetism of (Ga0.95,Fe0.05)Sb because its position in the IB model is different from that in the p-d Zener model. To determine the position of the Fe-3d IB in the VB experimentally, we have conducted resonant angle-resolved photoemission spectroscopy (r-ARPES) at the Fe L3 absorption edge. By using r-ARPES, the states derived from the orbitals which are relevant to the absorption process are resonantly enhanced, in our case the Fe d-states. Figure 5(a) shows the XAS spectrum at the Fe L3 edge of the present (Ga0.95,Fe0.05)Sb film decapped by annealing at around 300 oC. Its shape is similar to that of the reported XAS spectra of the (Ga,Fe)Sb thin films capped by an amorphous As layer35. Since the XAS spectra of the capped (Ga,Fe)Sb thin 10 films have been measured without annealing, this result indicates that the annealing hardly changes the 3d states of Fe in the (Ga0.95,Fe0.05)Sb film. A previous x-ray magnetic circular dichroism study of the capped (Ga,Fe)Sb thin films has demonstrated that the XAS peak of 708 eV originates from the ferromagnetic component35. The resonance enhancement of ARPES measured at this hν should therefore reflect the energy position of the ferromagnetic component of Fe-3d states in (Ga,Fe)Sb. Figures 5(b) and 5(c) show the r-ARPES images of the (Ga0.95,Fe0.05)Sb thin films taken at hν = 708 eV (on-resonance) and 704 eV (off-resonance) with p polarizations. A flat band appears in the on-resonance spectrum in the vicinity of EF in Fig. 5(b), while this band apparently disappears in the off-resonance spectrum in Fig. 5(c). The Fe-3d state in the vicinity of EF is consistent with the previous resonant photoemission spectroscopy (RPES) at the Fe L2,3 edge of (Ga,Fe)Sb35. The flat band observed with hν = 708 eV is therefore the Fe-3d IB, which is related to the ferromagnetism. Figure 5(d) shows the EDCs of the on- and off-resonance spectra from k// = 1.5 (k1) to k// = 2.1 (k2) in the reciprocal lattice unit of 2√2π/a. The difference of the on-resonance EDCs from the off-resonance ones [the blue areas between the pink and green dashed 11 curves in Fig. 5(d)] reveals the Fe-3d component of the ARPES spectra. Here, it should be noted here that the on-resonant signal of the IB is much enhanced compared with its real contribution to the density of states (DOS). Generally, the states induced by impurities doped in a single crystal are independent of the wave number owing to the random atomic distribution. In our case, however, the Fe-3d component areas strongly change with k|| along the VB dispersion as shown by the red dots in Figs. 5(b) and 5(d). This indicates that the intensity of this LH band is also enhanced at the Fe 2p-3d resonance through hybridization between the Fe-3d orbital and the ligand p band27. Therefore, the present observation provides experimental evidence that the Fe-3d IB hybridizes with the ligand p band. C. Discussion Finally, based on the experimental findings, we will discuss the band structure and the origin of the ferromagnetism of (Ga,Fe)Sb. From the observations described above, we have found that the Fe-3d IB located just above the VB maximum (VBM) crosses EF and hybridizes with the p band of GaSb, and that the LH, HH and SO bands of 12 (Ga0.95,Fe0.05)Sb similar to those of the host GaSb are located below EF, as shown in Fig. 6(a). Figure 6(b) shows a schematic diagram of the electronic structure of (Ga0.95,Fe0.05)Sb. The observation that the Fe-3d IB shows the Fermi cutoff indicates that the Fe-3d IB is partially occupied by electrons. Considering the result that the VBM is located below EF, it is probable that the concentration of hole carriers due to defects, if they exist, is much lower than that of the Fe ions. Additionally, if the Fe ions are ionic Fe3+, the majority(up)-spin states are fully occupied due to the d5 electronic configuration. It follows from these arguments that the number of d electrons in (Ga0.95,Fe0.05)Sb is expected to deviate from the half-filled d5 configuration through the p-d hybridization. The strength of the p-d hybridization is closely related with the symmetry of d electrons in (Ga,Fe)Sb. The five-fold degenerate state of Fe 3d splits into the two-fold degenerate states (e) and the three-fold degenerate states (t2) because the Fe ions of (Ga,Fe)Sb are in the tetrahedral crystal field. Due to the symmetries of the e and t2 states, while the e states do not hybridize with the ligand p bands, the t2 states hybridize with the p bands36. The p-d(t2) hybridization leads to the antibonding (t2a) and bonding (t2b) states. The p-d(t2) hybridization leads to the antibonding (t2a) and bonding (t2b) states, which have both the 13 Fe t2 and the ligand Sb p characters. Figure 6 shows the schematic energy diagram of (Ga,Fe)Sb based on the experimental findings. Here, ↑ and ↓ means majority(up) spin and minority(down) spin, respectively. For the majority-spin states, since the Fe-3𝑑↑ levels are located well below the VBM, the bonding 𝑡&'↑ and antibonding 𝑡&(↑ states have predominantly both the Fe t2 and Sb p characters, respectively. In contrast, since the Fe-3𝑑↓ levels are located above the VBM, the minority-spin bonding 𝑡&'↓ state is mainly composed of the ligand p orbitals, and the antibonding 𝑡&(↓ state primarily consists of the Fe t2 orbitals. Band widths of the bonding t2b and antibonding t2a states with the p characters will be comparable with or narrower than that of VB through the strong p-d(t2) hybridization due to the high covalency of the narrow-gap semiconductor GaSb. Considering that the VBM is located just below EF, it is probable that the majority-spin antibonding 𝑡&(↑ state crosses EF and is partially filled (see the left side of Fig. 6(c)) because the 𝑡&(↑ is located above the VBM. To keep the charge balance of Fe3+, the holes of the majority-spin 𝑡&(↑ state will be compensated by electron occupation of the minority-spin 𝑒↓ and/or 𝑡&(↓ states (see the right side of Fig. 6(c)). This is consistent with the first-principle calculations for (Ga,Fe)Sb that the majority-spin 𝑡&(↑ and the 14 minority-spin 𝑒↓ and/or 𝑡&(↓ states cross EF37,38. It follows from these arguments that the observed narrow Fe-3d IB crossing EF probably originates from the partially occupied minority-spin 𝑒↓ and/or 𝑡&(↓ states, and the observed p-d hybridized states in the vicinity of EF would be the 𝑡&(↑ and 𝑡&'↓ states with the primarily p characters. The present SX-ARPES study indicates that the electronic structure of (Ga0.95,Fe0.05)Sb is consistent with the IB model, where the Fe-3d IB crosses EF. This means that the carriers of (Ga,Fe)Sb are mainly derived from the 3d electrons even if the p-d hybridization is finite. It should be mentioned here that the transport properties of (Ga,Fe)Sb stay semiconducting when the Fe concentration is less than 20%28, although the observed Fe-3d IB of the (Ga0.95,Fe0.05)Sb thin film shows the Fermi cutoff. It is possible that either gap opening, or depletion of DOS near EF which is smaller than the experimental energy resolution occurs in the Fe-3d IB of (Ga,Fe)Sb. Additionally, the actual contribution of the Fe-3d IB to the transport properties is expected to be negligible because the Fermi cutoff of the Fe-3d IB (or the spectral weight at EF) is fairly small in the off-resonance spectra as shown in Fig. 5(b). The DOS of the Fe-3d IB near EF will increase with increasing the Fe concentration, resulting in the increase of the conductivity 15 and TC of (Ga,Fe)Sb with higher Fe concentrations. It follows from these arguments that the double-exchange interaction in the Fe-3d IB crossing EF is the origin of the ferromagnetism in (Ga,Fe)Sb. It has been reported that the local Fe concentration in (Ga,Fe)Sb fluctuates on the several-nm scale, and zinc-blende Fe-rich clusters are formed in the (Ga,Fe)Sb matrix [10, 22]. Since our (Ga,Fe)Sb thin film with x = 0.05 is expected to have a less fluctuation of the local Fe concentration, further studies including a systemic SX-ARPES measurements of the VB structure of (Ga,Fe)Sb with varying x is necessary to clarify the origin of the magnetic interaction. In addition, not only the conventional valence of Fe but also the intermediate valence of Fe should be taken into account because of the p-d hybridization. To identify the local electronic structure of Fe ions including the electronic structure parameters such as the charge-transfer energy and strength of Coulomb interaction, it is indispensable to conduct other measurements sensitive to the local electronic structure, e.g., resonant inelastic x-ray scattering (RIXS) combined with cluster-model calculation. The effectiveness of RIXS for FMS has been confirmed in (Ga,Mn)As39. 16 IV . SUMMARY We have performed SX-ARPES measurements on a (Ga0.95,Fe0.05)Sb thin film to obtain the information about its VB structure, the location of the Fe-3d IB, and its hybridization with the p band. By capping samples with an amorphous Sb layer and removing the cap by annealing in vacuo just before performing measurements, we have succeeded in observing the 3D bulk VB structure in the samples with clean surfaces. Experimentally, the band dispersion of (Ga0.95,Fe0.05)Sb is similar to that of GaSb. This indicates that the Fe ions hardly affect the band dispersion of the host GaSb. In addition, the non-dispersive Fe-3d IB hybridized with the ligand p bands is located around EF, indicating that the carriers of (Ga,Fe)Sb have the d-like character. Based on our results, the electronic structure of (Ga,Fe)Sb is consistent with the IB model. It is probable that the partially filled Fe-3d IB is composed of the minority-spin 𝑒↓ and/or 𝑡&(↓ states through the p-d hybridization. Thus, double-exchange interaction between Fe ions would be the origin of the ferromagnetic interaction of (Ga,Fe)Sb. To thoroughly study the ferromagnetic mechanism of (Ga,Fe)Sb with higher Fe concentrations, systematic studies 17 of the VB structure and the local electronic state of Fe-3d electrons on the (Ga,Fe)Sb thin films with various Fe concentrations are needed. ACKNOWLEWDGMENTS This work was supported by a Grant-in-Aid for Scientific Research (Grant Nos. 15H02109, 16H02095, 17H04922, 18H05345, and 23000010), Core-to-Core Program A. Advanced Research Networks from JSPS, and CREST of JST (No. JPMJCR1777), Japan. This work was partially supported the Spintronics Research Network of Japan (Spin-RNJ). Supporting experiments at SPring-8 were approved by the Japan Synchrotron Radiation Research Institute (JASRI) Proposal Review Committee (Proposal No. 2018A3841 and 2019A3841). REFERENCES 1 T. Dietl, A. Haury, and Y. Merle d’Aubigné, Phys. Rev. B 55, R3347 (1997). 2 H. Munekata, H. Ohno, S. von Molnar, A. Segmüller, L.L. Chang, and L. Esaki, Phys. Rev. Lett. 63, 1849 (1989). 3 H. Ohno, H. Munekata, S. von Molnár, and L.L. Chang, J. Appl. Phys. 69, 6103 (1991). 4 H. Munekata, H. Ohno, R.R. Ruf, R.J. Gambino, and L.L. Chang, J. Cryst. Growth 111, 1011 (1991). 5 H. Ohno, H. Munekata, T. Penney, S. von Molnár, and L.L. Chang, Phys. Rev. Lett. 68, 2664 (1992). 6 H. Ohno, A. Shen, F. Matsukura, A. Oiwa, A. Endo, S. Katsumoto, and Y. Iye, Appl. Phys. Lett. 69, 363 (1996). 7 A. Van Esch, L. Van Bockstal, J. De Boeck, G. Verbanck, A.S. van Steenbergen, P.J. Wellmann, B. Grietens, R. Bogaerts, F. Herlach, and G. Borghs, Phys. Rev. B 56, 13103 (1997). 8 T. Hayashi, M. Tanaka, T. Nishinaga, H. Shimada, H. Tsuchiya, and Y. Otuka, J. Cryst. Growth 175–176, 1063 (1997). 9 Y. Ohno, D.K. Young, B. Beschoten, F. Matsukura, H. Ohno, and D.D. Awschalom, Nature 402, 790 (1999). 10 M. Tanaka and Y. Higo, Phys. Rev. Lett. 87, 26602 (2001). 11 L. Chen, X. Yang, F. Yang, J. Zhao, J. Misuraca, P. Xiong, and S. von Molnár, Nano 18 Lett. 11, 2584 (2011). 12 T. Schallenberg and H. Munekata, Appl. Phys. Lett. 89, 42507 (2006). 13 P. Nam Hai, L. Duc Anh, S. Mohan, T. Tamegai, M. Kodzuka, T. Ohkubo, K. Hono, and M. Tanaka, Appl. Phys. Lett. 101, 182403 (2012). 14 A. V Kudrin, Y.A. Danilov, V.P. Lesnikov, M. V Dorokhin, O. V Vikhrova, D.A. Pavlov, Y. V Usov, I.N. Antonov, R.N. Kriukov, A. V Alaferdov, and N.A. Sobolev, J. Appl. Phys. 122, 183901 (2017). 15 N.T. Tu, P.N. Hai, L.D. Anh, and M. Tanaka, Appl. Phys. Lett. 105, 132402 (2014). 16 N.T. Tu, P.N. Hai, L.D. Anh, and M. Tanaka, Appl. Phys. Express 11, 63005 (2018). 17 N.T. Tu, P.N. Hai, L.D. Anh, and M. Tanaka, Appl. Phys. Express 12, 103004 (2019). 18 S. Goel, L.D. Anh, S. Ohya, and M. Tanaka, Phys. Rev. B 99, 14431 (2019). 19 T. Dietl, Nat. Mater. 9, 965 (2010). 20 T. Dietl, H. Ohno, F. Matsukura, J. Cibert, and D. Ferrand, Science (80-. ). 287, 1019 (2000). 21 T. Dietl, H. Ohno, and F. Matsukura, Phys. Rev. B 63, 195205 (2001). 22 M. Berciu and R.N. Bhatt, Phys. Rev. Lett. 87, 107203 (2001). 23 S. Ohya, I. Muneta, Y. Xin, K. Takata, and M. Tanaka, Phys. Rev. B 86, 94418 (2012). 24 S. Souma, L. Chen, R. Oszwałdowski, T. Sato, F. Matsukura, T. Dietl, H. Ohno, and T. Takahashi, Sci. Rep. 6, 27266 EP (2016). 25 J. Okabayashi, A. Kimura, O. Rader, T. Mizokawa, A. Fujimori, T. Hayashi, and M. Tanaka, Phys. Rev. B 64, 125304 (2001). 26 A.X. Gray, J. Minár, S. Ueda, P.R. Stone, Y. Yamashita, J. Fujii, J. Braun, L. Plucinski, C.M. Schneider, G. Panaccione, H. Ebert, O.D. Dubon, K. Kobayashi, and C.S. Fadley, Nat. Mater. 11, 957 EP (2012). 27 M. Kobayashi, I. Muneta, Y. Takeda, Y. Harada, A. Fujimori, J. Krempaský, T. Schmitt, S. Ohya, M. Tanaka, M. Oshima, and V.N. Strocov, Phys. Rev. B 89, 205204 (2014). 28 N.T. Tu, P.N. Hai, L.D. Anh, and M. Tanaka, Phys. Rev. B 92, 144403 (2015). 29 V.N. Strocov, T. Schmitt, U. Flechsig, T. Schmidt, A. Imhof, Q. Chen, J. Raabe, R. Betemps, D. Zimoch, J. Krempasky, X. Wang, M. Grioni, A. Piazzalunga, and L. Patthey, J. Synchrotron Radiat. 17, 631 (2010). 30 V.N. Strocov, X. Wang, M. Shi, M. Kobayashi, J. Krempasky, C. Hess, T. Schmitt, and L. Patthey, J. Synchrotron Radiat. 21, 32 (2014). 31 A. Damascelli, Z. Hussain, and Z.-X. Shen, Rev. Mod. Phys. 75, 473 (2003). 32 I. Vurgaftman, J.R. Meyer, and L.R. Ram-Mohan, J. Appl. Phys. 89, 5815 (2001). 33 K. Sriharsha, L.D. Anh, N.T. Tu, S. Goel, and M. Tanaka, APL Mater. 7, 21105 (2019). 19 34 N.T. Tu, P.N. Hai, L.D. Anh, and M. Tanaka, Appl. Phys. Lett. 108, 192401 (2016). 35 S. Sakamoto, N.T. Tu, Y. Takeda, S.I. Fujimori, P.N. Hai, L.D. Anh, Y.K. Wakabayashi, G. Shibata, M. Horio, K. Ikeda, Y. Saitoh, H. Yamagami, M. Tanaka, and A. Fujimori, Phys. Rev. B 100, 35204 (2019). 36 Y. Zhang, X. Yuan, X. Sun, B.-C. Shih, P. Zhang, and W. Zhang, Phys. Rev. B 84, 75127 (2011). 37 H. Shinya, T. Fukushima, A. Masago, K. Sato, and H. Katayama-Yoshida, J. Appl. Phys. 124, 103902 (2018). 38 V.A. Gubanov, C.Y. Fong, and C. Boekema, Phys. Status Solidi 218, 599 (2000). 39 M. Kobayashi, H. Niwa, Y. Takeda, A. Fujimori, Y. Senba, H. Ohashi, A. Tanaka, S. Ohya, P.N. Hai, M. Tanaka, Y. Harada, and M. Oshima, Phys. Rev. Lett. 112, 107203 (2014). 20 FIG. 1. Sample structure of (Ga1-x,Fex)Sb thin films: (a) structure of the (Ga1-x,Fex)Sb (x = 0 and 0.05) thin films, (b) Brillouin zone (BZ) with the areas surrounded by red and purple lines measured in the out-of-plane and in-plane measurements, respectively. \n FIG. 2. Constant-energy mappings in the kz-k|| plane: (a)-(c) mappings for EB = 1.6 eV, 3.7 eV, and 7.0 eV, respectively and red solid lines represent the BZ in Fig. 1(b). \nkz (2π/a)-10k|| (2√2π/a)-10-10141315(a) EB = 1.6 eV(b) EB = 3.7 eV(c) EB = 7.0 eVKXΓ\nLowHigh 21 \n FIG. 3: Band dispersion around the Γ point of a (Ga0.95,Fe0.05)Sb thin film: (a) FSM in the kz-k|| plane where the yellow curve represents the cut at hν = 885 eV, and the red solid lines represent the BZ; (b) FSM in the k[-110] - k[-1-10] plane taken at hν = 885 eV where purple solid lines represent the BZ; (c), (d) ARPES images along the Γ-K-X line taken with p and s polarizations, respectively. The spectra are measured at hν = 885 eV, and the LH, HH, and SO denote the light-hole, heavy-hole and split-off bands, respectively; (e) EDCs along the Γ-K-X line corresponding to the ARPES spectrum shown in panel (c), and the blue vertical bars are a guide to the eyes tracing the band dispersion of the LH band. \n86420Binding Energy (eV)\n86420Binding Energy (eV)10k|| (2√2π/a)-101kz (2π/a)-10k|| (2√2π/a)\n-101k[-1-10] (2√2π/a)10k[-110] (2√2π/a)ΓKX\n p-pol.885 eV\n s-pol.885 eVHHLHSOΓKX\nΓKXEB = 0 eV\nEB = 0 eV885 eV 708 eV 704 eV10151314ΓKX(a)\n(b)(c)\n(d)\nLowHighIntensity (a.u.)420Binding Energy (eV) EDC, p-pol. 885 eV\nΓKX(e) 22 \n FIG. 4. Band dispersion near the Γ point of GaSb; (a) ARPES image of GaSb along the Γ-K-X line measured with p polarization at hν = 880 eV, (b) EDCs of the (Ga0.95,Fe0.05)Sb (red curves, from Fig. 3(c)) and GaSb (blue curves, from Fig. 4(a)) thin films. \n86420Binding Energy (eV)10.50k|| (2√2π/a)86420Binding Energy (eV)Intensity (a.u.)ΓKX\n p-pol. 880 eVLHSOGaSbGa0.95Fe0.05SbGaSb\nΓKX(a)(b)\nLowHigh 23 \n FIG. 5. Resonant ARPES spectra of (Ga0.95,Fe0.05)Sb: (a) Fe L3 XAS spectrum for hv = 708 eV and 704 eV are for the on-resonance and off-resonance, respectively; (b), (c) on- and off-resonant ARPES images measured with p polarization, respectively; (d) EDCs of the on-resonant ARPES (pink solid curves, from Fig. 5(b)) and off-resonant ARPES (green dashed curves, from Fig. 5(c)) between k|| =1.5 (k1) and k|| = 2.1 (k2). Shaded (blue) area denotes the Fe-3d component of ARPES spectra. Red dots on pink solid curves in Figs. 5(b) and 5(d) represent the points where the on-resonant intensity is remarkably stronger than the off-resonant intensity. \n715710705700Photon Energy (eV)\n43210Binding Energy (eV)210k|| (2√2π/a)210k|| (2√2π/a)Intensity (a.u.)\nIntensity (a.u.)10Binding Energy (eV)(b)p-pol. 708 eV (on)IB(c)p-pol. 704 eV (off)Ga0.95Fe0.05Sb Fe L3 XASonoff on off\nk2k1(a)(d)\nLowHigh 24 \n FIG. 6: Electronic structure of (Ga0.95,Fe0.05)Sb: (a) ARPES image from Fig.3(c) with added Fe-3d IB in the vicinity of EF of Fig. 5(b). (b) Schematic diagram for the density of states (DOS) of (Ga0.95,Fe0.05)Sb. (c) Schematic energy diagram of (Ga0.95,Fe0.05)Sb where the p-d hybridization splits t2 states into antibonding (t2a) and bonding (t2b) states. CB means the conduction band. \n" }, { "title": "2004.03784v1.Charge_to_spin_conversion_efficiency_in_ferromagnetic_nanowires_by_spin_torque_ferromagnetic_resonance__Reconciling_lineshape_and_linewidth_analysis_methods.pdf", "content": "arXiv:2004.03784v1 [cond-mat.mes-hall] 8 Apr 2020Charge-to-spin conversion efficiency in ferromagnetic nano wires by spin torque\nferromagnetic resonance: Reconciling lineshape and linew idth analysis methods\nJun-Wen Xu and Andrew D. Kent∗\nCenter for Quantum Phenomena, Department of Physics, New Yo rk University, New York 10003, USA\n(Dated: April 9, 2020)\nSpinorbit torquesare ofgreat interest for switching thema gnetization direction innanostructures,\nmoving skyrmions and exciting spin waves. The standard meth od of determining their efficiency\nis by spin torque ferromagnetic resonance (ST-FMR), a techn ique that involves analyzing the res-\nonance linewidth or lineshape. On microstuctures these two analysis methods are quite consistent.\nHerewe presentST-FMR results onpermalloy (Ni 80Fe20)nanowires —with widths varyingfrom 150\nto 800nm — that show that the standard model used to analyze th e resonance linewidth and line-\nshape give different results; the efficiency appears greatly e nhanced in nanowires when the lineshape\nmethod is used. A ST-FMR model that properly accounts for the sample shape is presented and\nshows much better consistency between the two methods. Micr omagnetic simulations are used to\nverify the model. These results and the more accurate nanowi re model presented are of importance\nfor characterizing and optimizing charge-to-spin convers ion efficiencies in nanostructures.\nI. INTRODUCTION\nSpin orbit torques (SOTs) are being actively consid-\nered for using in the next generation memory devices\nfor magnetization switching [1, 2], spin oscillators [3, 4]\nand racetrack memories, including those using magnetic\nskyrmions [5–8]. SOTs are fundamentally based on\ncharge-to-spin conversion by the spin-Hall effect [9] or\nRashba effect [10]. These effects result in a spin cur-\nrent or spin accumulation that is transverse to the di-\nrection of charge current flow. An advantage of SOTs\nis that a charge current does not need to flow through\nthe magnetic layer to switch its magnetization direction.\nThe torques can thus be used to switch magnetic insula-\ntors[11] orthe freelayerofamagnetictunnel junction [1]\nwithout current flowing through the insulating layer.\nSOTsareprincipallyofinterestinnanostructuredsam-\nples, samples with minimum dimension less than a mi-\ncron. In such samples the switching current density can\nbe relatively large while the current is still small and the\ntorques on the magnetization associated with chargecur-\nrent induced Oersted fields can be much smaller than the\nSOT. However, the magnitude and form of the torques\nare most often characterized in micron scale samples.\nThe most widely used SOT characterization method is\nspin torque ferromagnetic resonance (ST-FMR) [12, 13].\nIn this technique a radio frequency (rf) charge current\nleads to a rf spin torque that excites magnetization dy-\nnamics, with the largest response occurring at the fer-\nromagnetic resonance (FMR) frequency. The magni-\ntude of the torque can be determined by either ana-\nlyzing the linewidth of the response or the lineshape\nof the resonance [14, 15]. This allows extracting the\neffective charge-to-spin conversion efficiency, which is\nproportional to the spin-Hall angle θSH=js/jc[16].\nThe linewidth and lineshape analysis methods have been\n∗andy.kent@nyu.edushown to give consistent results when applied to mi-\ncrostructures [15, 17–19].\nIn this paper we show that when ST-FMR results on\nnanostructures are analyzed with the standard ST-FMR\nmodel the results are generally not consistent; with the\nlineshape analysis method the efficiency appears greatly\nenhanced in nanowires. We present ST-FMR results on\npermalloy nanowires with linewidths varying from 150\nto 800nm and a model that properly accounts for the\nsample geometry. Our nanowire model provides much\nbetter consistency between the lineshape and linewidth\nanalysis methods.\nII. EXPERIMENT\nExperiments were conducted on Permalloy (Py,\nNi80Fe20) nanowires fabricated from thin films. The\nfilms are grownby dc magnetron sputtering on thermally\noxidized silicon substrates with layer stack SiO 2/Ta\n(3)/Py(5)/Pt(6), where the numbers are the layer thick-\nnesses in nm. Ta is a seed layer [20] and the Pt layer is\nthe main source of the spin current. It also protects the\nPy from oxidation. Electron beam lithography followed\nby argon ion milling is used to define the sample geome-\ntry. We deposit Ti(5)/Au(50) by evaporation and liftoff\nfor the contact pads. An image of the nanowire and con-\ntact pads is shown in the inset of Fig. 1. The width of\nnanowires varies from 150 to 800nm with a fixed aspect\nratio of 20.\nFigure1shows the ST-FMR setup. The nanowire is\noriented at 45 degrees to the applied in-plane field in or-\nderto havea largeST-FMR signal. Arfchargecurrentis\napplied to the sample using a ground-signal (GS) probe\nconnected to contact pads to the nanowire. The rf cur-\nrent is input to the high frequency port of a bias tee,\nwhile the bias tee’s low frequency port is connected to a\ndc current source and a lock-in amplifier.\nThe measurement principal is as follows. The rf cur-\nrent produces an rf torque on the ferromagnetic layer,2\nFigure1. ST-FMRsetup. Thenanowire isoriented45degrees\nto the external field B. An rf and dc current are applied using\na bias tee, while a small field coil is used to modulate the\napplied field at a low frequency.\ncausingthis layer’smagnetizationto precess. The largest\nprecession amplitude occurs at the FMR frequency. The\noscillation in the magnetization leads to oscillations in\nthe nanowire resistance due to the anisotropic magne-\ntoresistance (AMR) of Py. A dc voltage Vmixappears\nacross the nanowire due to the mixing of the rf charge\ncurrent and resistance oscillations at the same frequency.\nIn order to increase the signal-to-noise ratio, a small\namplitude magnetic field of 0 .2mT (an amplitude much\nless than the FMR linewidth) is modulated at low fre-\nquency 727Hz, a frequency far less than the FMR fre-\nquency. The lock-in amplifier is set to detect the signal\natthis lowfrequency. Themodulated voltagesignalmea-\nsured isV(B)∝dVmix/dB[21]. A single measurement\ncorresponds to sweeping the external field from high to\nlow values at fixed rf current amplitude and frequency.\nThe sweep is from high to low field to ensure each mea-\nsurements starts from a saturated magnetic state. For\neach rf frequency, the field is swept multiple times and\nthe signal is averaged in order to further increase the\nsignal-to-noise ratio. The rf frequency is varied from 9\nto 15GHz and magnetic fields up to 0 .3T are applied.\nThe power of our rf source is fixed at 10dBm and we\nhave verified that the resonance response is in the linear\nregime.\nThe lineshape analysis method consists of analyzing\nthepeakshape. Thepeakinthelock-involtageasafunc-\ntion of magnetic field, V(B), can be decomposed into the\nsum of the derivative of Lorentzian and anti-Lorentzian\nfunctions (see the supplementary materials for further\ndetails):\nV(B) =−S(B−B0)∆+A/bracketleftbig\n(B−B0)2−(∆/2)2/bracketrightbig\n[(B−B0)2+(∆/2)2]2,(1)\nwhereB0is the resonance field, Bis the external field\nin vacuum (i.e. B=µ0H, whereµ0is the permeability\nof free space) and ∆ is the resonance full width at half\nmaximum (FWHM). SandAare Lorentzian and anti-\nLorentzian amplitudes, respectively. In contrast, for the\nlinewidth analysis method, a dc current is applied to thesample through the low frequency port of the bias tee.\nThe variation of the resonance peak linewidth with dc\ncurrent is used to determine the efficiency.\nIn the following we denote the standard ST-FMR\nmodel the thin film model, as this model assumes the fer-\nromagnethasaneasy-planemagneticanisotropy,withno\npreferred magnetization axis in the plane. The nanowire\nmodel we present considers the in-plane magnetic shape\nanisotropy that we discuss further below.\nA. Results and analysis with thin film model\nFigure2(a)-(c) show spectra of 150, 400 and 800nm\nlinewidth nanowires at a fixed rf frequency of 12GHz,\nwhere each spectra is normalized to its maximum value.\nIt is clear that the signal-to-noise ratio is more than ade-\nquate for detailed analysis. We fit the spectra to Eq. 1to\ndetermine the Lorentzian Sand anti-Lorentzian Aam-\nplitudes, the resonance field B0and the linewidth ∆.\nIn the thin film model the ratio of the Lorentzian\nand anti-Lorentzian amplitudes is used to compute the\ncharge-to-spin conversion efficiency, ξ, given by [15]:\nξ=S\nA/radicalbigg\n1+µ0Meff\nB0eµ0Mstd\n/planckover2pi1, (2)\nwheretis the thickness of ferromagnetic layer and dis\nthe thickness of heavy metal layer. Meffis the effective\nmagnetization that characterizes the ferromagnet’s easy\nplane anisotropy and Msis its saturation magnetization,\neis the electron charge, and /planckover2pi1is the reduced Planck’s\nconstant. We note that this model assumes a negligible\nfield like SOT; it assumes that the field-induced torque is\nassociatedwiththe Oerstedfield fromthe chargecurrent.\nThis is a reasonableassumption for our samples, as it has\nbeen reported that the field-like SOT decays with mag-\nnetic layer thickness, with a characteristic decay length\nof≃2nm [22]. Thus for our 5nm thick ferromagnetic\nlayer this torque will be far less than the anti-damping\ntorque and the Oersted field induced torque.\nUsing the fitting parameters derived from the data in\nFig.2andtakingMeff=Ms, Eq.2isusedtodetermine ξ.\nThe result is shown in Fig. 2(d). The efficiency increases\ndramatically (i.e. by more than a factor of 2) as the wire\nlinewidth is reduced. For larger nanowire widths ( /greaterorsimilar400\nnm),ξ= 0.65 which is consistent with that reported\nvalue for Pt/Py interfaces [15, 17, 18]. It is reasonable\nto associate the main component of SOT with this in-\nterface because Ta is much more resistive [23] than Pt\nand thinner in our samples, so most of the current passes\nthrough the Pt layer.\nWe also studied the efficiency using the linewidth\nanalysis method, expecting to see a similar trend with\nnanowire width. As noted above, in this method a dc\ncurrent is applied to the nanowire and the ST-FMR\nlinewidth is determined as a function of the current. For\none polarity ofthe current the SOT opposesthe damping\nand leads to a reduced ST-FMR linewidth, while for the3\n00.51.0Data\nLorentzian\nanti-Lorentzian\nSum\n200 400 600 80000.050.100.15\nWidth (nm)ξ00.51.0\n0.1 0.2 0.300.51.0\nB (T)Normalized Voltage\n(d)(a) 150 nm\n(b) 400 nm\n(c) 800 nm\nFigure 2. (a)-(c) Normalized spectra of 150, 400 and 800 nm\nwidth nanowires at 12 GHz. The normalization voltages are\n2.43µV, 1.71µV and 1.48µV respectively. The red and green\ncurves show the anti-Lorentzian and Lorentzian components\nof the resonance peak respectively. (d) The efficiency versus\nnanowire width determined from the lineshape using the thin\nfilm model.\noppositecurrent,SOTincreasesthedampingandtheST-\nFMR linewidth increases [24]. Figure 3shows selected\nspectra at fixed rf frequency for several dc bias currents.\nFitting these spectra to Eq. 1we determine the linewidth\nas function of the dc bias. The results are shown in\nFig.3(b). For positive field, the linewidth decreases with\nincreasing bias current, and vice versa. The slope of the\nlinewidthversusdcbiascurvesis −2.2×10−3mT/µAfor\npositive field and 2 .3×10−3mT/µA for negative field.\nIn the thin film model the efficiency is related to the\nslope of the ST-FMR resonance linewidth versus dc bias\ncurrent [14, 15]:\nξ=γe(B0+µ0Meff/2)Mst\n2π/planckover2pi1fsinφd∆\ndjc, (3)\nwherejcisthe currentdensity in the Pt layerand φis the\nanglebetweenchargecurrentandexternalmagneticfield.\nFigure3(c) shows the resulting ξversus nanowire width;\nξvaries between 0.062 - 0.077. Comparing Fig. 2(d) and\nFig.3(c) it is clear there is a significant discrepancy in\nthe SOT efficiencies deduced from these models. It is\nalso clear that some basic physics in the modeling of the\nefficiency in nanowires is not captured by the standard\nthin film model.0.15 0.16 0.17−1012\nB (T)V (µV)400\n160\n0\n-160\n-400\n−400 0 400192021\nCurrent (µA)∆ (mT)Positive ®eld\nNegative ®eld\n200 400 600 80000.030.060.09\nWidth (nm)ξ(b)\n(c)(a) I (µA)\nFigure 3. (a) ST-FMR spectra of a 300nm width nanowire\nat 12GHz at several dc bias currents. (b) Linewidth as a\nfunction of dc current for both positive and negative field\npolarities. (c) The efficiency as a function of the nanowire\nwidth determined using the linewidth analysis method.\nB. Nanowire model\nWe begin by analyzing the dependence of resonance\nfieldon rffrequencytotest abasicassumptionofthe thin\nfilm model. The model assumes an easy-plane magnetic\nanisotropy with a resonance frequency-field relation:\nf2\nB0=/parenleftBigγ\n2π/parenrightBig2\n(B0+µ0Meff). (4)\nThat isf2/B0should be a straight line when plotted\nversus field B0, with a slope proportional to the gy-\nromagnetic ratio squared and intercept proportional to\nMeff. The experimental data are shown in Fig. 4. It is\nclearthat the 800nm data follows the expectations of the\nmodel. However, the data for the 150nm width wire de-\nviatesstronglyfromthestraightlinetrend(orangecurve)\nfor resonance fields less than about 0 .15T.\nAssumingthatappliedfieldsgreaterthan0 .15Tarere-\nquired to fully saturate the magnetization of the 150nm\nwidth wire, one may omit the lower resonance field data\nfrom the fits to Eq. 4(see the orange line in Fig. 4).\nFollowing this approach for all samples we find the fit-\nting parameters Meffandγ/(2π) given in Table I. For\nwide nanowires the parameters are those expected for4\nWidth (nm) 150 200 300 400 600 800\nµ0Meff(T) 1.35(5) 1.19(1) 1.19(2) 1.11(2) 1.04(1) 0.96(1)\nγ/2π(GHz/T)24.7(4) 26.2(1) 26.3(2) 27.2(2) 28.0(1) 28.9(1)\nTable I. Effective magnetization and gyromagnetic ratio fro m\nfits to Eq. 4 for different width nanowires.\nPy,µ0Meff≈µ0Ms= 1T and γ/(2π) = 28GHz /T.\nHowever, for the narrower samples, µ0Meffis larger and\nγ/(2π) is smaller. This does not make physical sense\nasµ0Meffis associated with the demagnetization field\nperpendicular to the film plane, which decreases as the\nwire width decreases. Without an additional easy plane\nmagnetic anisotropy (e.g. associated with spin-orbit in-\nteractions at the Py interfaces), Meff≤Ms. For exam-\nple, for the 150nm width nanowire, the demagnetization\nfield in the width direction ≃0.05T [25], which is not\nnegligible compared to the applied magnetic field in our\nexperiments (0.1 to 0 .3T).\nIt is thus clear that the in-plane magnetic anisotropy\nassociated with the sample shape is not negligible and\nneeds to be considered in the analysis of the resonance\nfield. The shape anisotropy can be described by a in-\nplane uniaxial anisotropy field BAparallel to the wire\naxis. The resulting resonance field condition becomes:\nf=γ\n2π/radicalbig\n(B0+BA)(B0+BA+µ0Meff).(5)\nThe additional field BAis associated with the average\ndemagnetization field in the width direction, i.e. the in-\nplane magnetic shape anisotropy. The fit to Eq. 5is\nshown as the green curve in Fig. 4. The fit accurately\ncaptures both the low and high field resonance data with\nthe fit parameters for all the nanowires shown in Ta-\nbleII. We now see, as expected, that Meffdecreases as\nthe wire width decreases. Further, the gyromagnetic ra-\ntio is nearly independent of wire width.\n0.10 0.15 0.20 0.258809109409701000\nB0 (T)f2/B0 (GHz2/T)\n150 nm data\n800 nm data\n150 nm Eq. 4\n150 nm Eq. 5\nFigure 4. Resonance field versus rf frequency for 150 and\n800nm nanowires, fit with Eq. 4 and Eq. 5Width (nm) 150 200 300 400 600 800\nµ0Meff(T) 0.80(8) 0.83(7) 0.94(2) 0.94(2) 0.96(4) 0.92(5)\nBA(mT) 14(3) 10(2) 7(1) 3(1) 1(1) 1(1)\nγ/2π(GHz/T)29.6(10) 29 .6(8) 28.7(2) 28.8(2) 28.9(5) 29.4(6)\nTable II. Effective magnetization, in-plane anisotropy fiel d\nand gyromagnetic ratio from fits to Eq. 5 for different width\nnanowires.\nAs a consequence of the shape anisotropy the demag-\nnetization field is not collinear with the applied field in\nour experiment. Thus the precession axis of the magne-\ntization in resonance is also no longer the applied field\ndirection. As a result the angle between the precession\naxis and the current is less than 45 degrees, φ<π/4.φ\ncan be calculated numerically by solving\nBsin(π/4) =Nxµ0Mssinφ+Btanφcos(π/4),(6)\nwhereNxis the demagnetization coefficient in the width\ndirection (see supplementary part 2).\nFurther, the charge current induced Oersted field for\nnanowires is different from the film case as well. The in-\nplane Oersted field decreases at the edge of the wire, as\nits direction becomes more out of the film plane. The\nST-FMR model assumes a uniform magnetization re-\nsponse, so the average in-plane Oersted field is consid-\nered. The average in-plane Oersted field for a very wide\nstrip (BOe=µ0Jd/2) is reduced, multiplied by a factor\n¯Θ(ǫ)<1 (see supplementary part 3):\n¯Θ(ǫ) =1\nπ/bracketleftbigg\n2arctanǫ−1\nǫln/parenleftbig\n1+ǫ2/parenrightbig/bracketrightbigg\n,(7)\nwhereǫ≡w/dis the ratio of width to the thickness of\nthe Pt layer. In the film case, ǫ→ ∞,¯Θ = 1. For a\n150nm nanowire, ǫ= 25 and ¯Θ≈0.89.\nIn the nanowire model lineshape analysis method:\nξ=S\nA/radicalbigg\n1+µ0Meff\nB0+BAeµ0Mstd\n/planckover2pi1¯Θ. (8)\nWhile in the linewidth analysis method, the expression\nis the same as that in Eq. 3. However, φnow needs to\nbe evaluated numerically using Eq. 6.\nFigure5shows the ST-FMR data analysis using our\nnanowire model. The efficiency ξusing both lineshape\nanalysis and linewidth analysis methods for all width of\nthe nanowires is between 0.6-0.7. Also, there is no longer\na significant enhancement in ξat small nanowire widths\nwith the lineshape analysis method.\nC. Nanowire lineshape versus linewidth analysis\nmethods\nWe find that the nanowire lineshape and linewidth\nanalysis methods are in reasonable accord with small\ndifferences, about 10 % differences in ξ, comparable to5\n00.050.100.15\nFilm model\nNanowire model\n200 400 600 80000.030.060.09\nWidth (nm)ξ\nFilm model\nNanowire model(a)\n(b)Lineshape analysis \nLinewidth analysis \nFigure 5. Comparison of the thin film and nanowire models\nusing (a) the lineshape analysis method and (b) the linewidt h\nanalysis method.\nthe measurement error. There are reasons that the line-\nshape method may be more accurate. First, this is be-\ncause the lineshape analysis method is self-calibrated;\nit is a measure of the ratio of the SOT to the Oersted\nfield torque which are both proportional to the current\ndensity in the heavy metal layer. Hence, the ratio of\nLorentzian and anti-Lorentzian peak amplitude is inde-\npendent of the charge current density. In contrast, in\nthe linewidth analysis method, it is important to ac-\ncurately estimate the current density passing through\nthe heavy metal layer. Second, in the linewidth anal-\nysis method, the angle between the magnetization pre-\ncession axis and charge current is estimated assuming a\nuniform demagnetization field for a cuboid shaped sam-\nple [25], which is an approximation (see the last part of\nSec.III). Third, during a field swept spectrum, this angle\nchanges with the external field. Here we take the angle,\nat the resonance field, as an approximation, which leads\nto an additional error. Finally, the most important set\nof data for the linewidth analysis method is the variation\nof the linewidth with the dc bias current. In our study,\nthis change is only 1mT when the current density varies\nby±2×1011A/m2, which is comparable to the 0 .2mT\nmodulation field. This introduces a systematic error in\ndeterminingthelinewidthvariationwithdccurrent. Fur-\nthermore, the external field control is of order 0 .01mT,\nlimiting the signal-to-noise ratio.\nIII. MICROMAGNETIC SIMULATIONS\nIn order to verify our nanowire model, we carried out\nmicromagnetic modeling using MuMax (see supplemen-\ntary part 4 for our code) [26, 27]. We simulate nanowires\nwith a width of 160, 320, 640 and 1280nm and an aspect\nratio of 16. As in the experiment, the magnetization isdriven by an oscillating Oersted field torque and SOT\nassociated with a charge current. After 4ns of simula-\ntion time, the spatial averaged magnetization amplitude\ntakes the form:\nmx(t) =mxcos(2πft−ψ), (9)\nwheremx(t) is the instantaneous xcomponent of mag-\nnetization, mxis the amplitude of oscillation and ψis\nthe phase delay between the magnetization response and\nrf current. We sweep the external field Bat a fixed rf\ncurrent amplitude and frequency. The ST-FMR signal\nresultsfrommixingoftherfcurrentandanisotropicmag-\nnetoresistance:\nVmix=1\nT/integraldisplay\nI(t)R(t)dt\n=−(IδRsinφcosφ)mxcosψ,(10)\nwhere the integral is over one period of the rf signal\nT= 1/fandmxcosψis a sum of a Lorentzian and anti-\nLorentzian function (see supplementary part 1). There-\nfore, we fit the simulation result mxcosψversusBusing\nthe expression\nmxcosψ=S(∆/2)+A(B−B0)\n(B−B0)2+(∆/2)2,(11)\nin order to extract the resonance field ( B0) the linewidth\n(∆), the Lorentzian amplitude ( S) and the anti-\nLorentzian amplitude ( A), as we do for the experimental\ndata.\nFitting the resulting resonance frequency versus mag-\nnetic field with Eq. 4(the thin film model) the trends of\nMeffandγ/(2π) versus wire width are similar to those\nfound for the experimental data (c.f. Table Iand sup-\nplementary part 5 Table 1 for the simulation results);\nMeffincreases with decreasing wire width. Since in the\nsimulations µ0Ms= 1T andγ/(2π) = 28.0GHz/T, this\ndiscrepancy is unphysical. However, using the nanowire\nexpression, Eq. 5, gives reasonable values of Meffand\nγ/(2π) (see supplementary part 5 Table 2 for the simu-\nlation results).\nThe efficiency extracted from the models is plotted in\nFig.6. Note that this should be compared to the ef-\nficiency input into the micromagnetic simulations, ξ=\n0.05. In the wide wire limit, both the thin film model\nandnanowiremodelareingoodagreementusingtheline-\nshapeanalysisandlinewidth analysismethods. However,\nwhenthewirewidthbecomesnarrower,thethinfilmline-\nshape analysis gives a ξthat increases (Fig. 6), as seen\nin experiment (Fig. 2(d)). Using the nanowire model ξ\nis closer to 0.05. The origin of the enhancement of ξ\nis now clear. It is associated with the thin film model\noverestimating Meffand the Oersted field, ¯Θ.\nIn the linewidth analysis method, the thin film model\nshows that ξdecreases in narrower wires. But in the\nnanowire model, there is no longer such a decrease. We\nhave not seen such a decrease in our experimental re-\nsults, Fig. 3(c). This is because in the thin film model6\n0 500 1000 15000.040.050.060.07\nWidth (nm)ξFilm model Film model\nNanowire model Nanowire model\nInput valueLineshape analysis Linewidth analysis\nFigure 6. Simulation results of efficiency found using line-\nshape and linewidth analysis methods with both the thin film\nand nanowire models with for different wire widths. The\ndashed curve is the value of the efficiency used in the sim-\nulation, ξ= 0.05.\n(Eq.3), we overestimate Meffwhile we underestimate\n1/sinφ, which compensate each another.\nFinally, as discussed near Eq. 6, the equilibrium mag-\nnetization angle φwas calculated with demagnetiza-\ntion coefficients associated with a uniformly magnetized\ncuboid. To estimate the error associated with this as-\nsumption, we compared the average magnetization an-\ngle determined in micromagnetics simulations from that\nfound assuming a uniformly magnetized cuboid. For the\nlatter, the demagnetization field is calculated following\nRef. [25] and input into a micromagnetics simulation as\na fixed uniform field (turning off the internal demagneti-\nzation field in the simulation, see supplementary part 6\nfor details). The relation of Band sinφboth using mi-cromagnetics and a uniform demagnetization field with\ndifferent width of nanowires is plotted in supplementary\npart6. Wefindthatassumingauniformdemagnetization\nand following our procedure using Eq. 6produces a neg-\nligible error in the nanowire linewidth analysis method.\nIV. SUMMARY\nIn summary we have introduced an analytic ST-FMR\nmodel for nanowires that accounts for their shape in\na straightforward way. The model gives reliable effi-\nciency results either by analyzing the ST-FMR lineshape\nor the ST-FMR linewidth as a function of current. As\nthe primary interest in spin orbit torques is in exciting\nmagnetization dynamics and switching the magnetiza-\ntion of nanostructures, our model can be of importance\nin reliablecharacterizationandoptimizing charge-to-spin\nconversion efficiencies in such structures. Further, the\nmodel can be extended to other sample shapes and to\ninclude additional physics associated, for example, with\nthe quantization of spin wave modes in nanostructures\nand spin-pumping.\nACKNOWLEDGMENTS\nWe thank Dr. Christopher Safranski for discussions of\nST-FMR analysis methods and Dr. Nahuel Statuto for\nthe guidance in the micromagnetic modeling. This re-\nsearchwas supported by DARPA Grant No. D18AP0000\nand the National Science Foundation under Grant No.\nDMR-1610416. The nanostructures were realized at the\nAdvanced Science Research Center NanoFabrication Fa-\ncility of the Graduate Center at the City University of\nNew York.\n[1]L. Liu, C. F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and\nR. A. Buhrman, Science 336, 555 (2012).\n[2]I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten,\nM. V. Costache, S. Auffret, S. Bandiera, B. Rodmacq,\nA. Schuhl, and P. Gambardella, Nature 476, 189 (2011).\n[3]V. E. Demidov, S. Urazhdin, and S. O. Demokritov,\nNature Materials 11, 1028 (2012) .\n[4]Z. Duan, A. Smith, L. Yang, B. Youngblood, J. Lindner,\nV. E. Demidov, S. O. Demokritov, and I. N. Krivorotov,\nNature Communications 5, 5616 (2014) .\n[5]A. Fert, V. Cros, and J. Sampaio, Nature Nanotechnol-\nogy8, 152 (2013).\n[6]S. Woo, K. M. Song, H.-S. Han, M.-S. Jung, M.-Y.\nIm, K.-S. Lee, K. S. Song, P. Fischer, J.-I. Hong,\nJ. W. Choi, B.-C. Min, H. C. Koo, and J. Chang,\nNature Communications 8, 15573 (2017) .\n[7]S. A. Montoya, R. Tolley, I. Gilbert, S.-\nG. Je, M.-Y. Im, and E. E. Fullerton,Physical Review B 98, 104432 (2018) .\n[8]W. Jiang, X.Zhang, G. Yu, W. Zhang, X.Wang, M. Ben-\njamin Jungfleisch, J. E. Pearson, X. Cheng, O. Heinonen,\nK. L. Wang, Y. Zhou, A. Hoffmann, and S. G. E.\nte Velthuis, Nature Physics 13, 162 (2016) .\n[9]A.Hoffmann, IEEE Transactions on Magnetics 49, 5172 (2013) .\n[10]A. Manchon and S. Zhang, Phys. Rev. B 78, 212405\n(2008).\n[11]P. Li, T. Liu, H. Chang, A. Kalitsov, W. Zhang,\nG. Csaba, W. Li, D. Richardson, A. DeMann, G. Ri-\nmal, H. Dey, J. S. Jiang, W. Porod, S. B. Field, J. Tang,\nM. C. Marconi, A. Hoffmann, O. Mryasov, and M. Wu,\nNature Communications 7, 12688 (2016) .\n[12]J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N.\nKrivorotov, R. A. Buhrman, and D. C. Ralph, Phys.\nRev. Lett. 96, 227601 (2006).\n[13]A. A. Tulapurkar, Y. Suzuki, A. Fukushima, H. Kub-\nota, H. Maehara, K. Tsunekawa, D. D. Djayaprawira,7\nN. Watanabe, and S. Yuasa, Nature438, 339 (2005) .\n[14]S. Petit, C. Baraduc, C. Thirion, U. Ebels,\nY. Liu, M. Li, P. Wang, and B. Dieny,\nPhysical Review Letters 98, 077203 (2007) .\n[15]L. Liu, T. Moriyama, D. C. Ralph, and R. A. Buhrman,\nPhysical Review Letters 106, 036601 (2011) .\n[16]ST-FMR experiments relate the SOT to the charge cur-\nrent. The SOT depends on the spin current as well as\nother factors such as the interface transparency and the\nspin diffusion lengths in the materials.\n[17]W. Zhang, W. Han, X. Jiang, S.-H. Yang, and S. S. P.\nParkin,Nature Physics 11, 496 (2015) .\n[18]N. Reynolds, P. Jadaun, J. T. Heron, C. L. Jermain,\nJ. Gibbons, R. Collette, R. A. Buhrman, D. G. Schlom,\nand D. C. Ralph, Physical Review B 95, 064412 (2017) .\n[19]A. Ganguly, K. Kondou, H. Sukegawa, S. Mitani,\nS. Kasai, Y. Niimi, Y. Otani, and A. Barman,Applied Physics Letters 104, 072405 (2014) .\n[20]M.-H. Nguyen, D. C. Ralph, and R. A. Buhrman,\nPhysical Review Letters 116, 126601 (2015) .\n[21]A. M. Gonalves, I. Barsukov, Y.-J. Chen,\nL. Yang, J. A. Katine, and I. N. Krivorotov,\nApplied Physics Letters 103, 172406 (2013) .\n[22]C.-F. Pai, Y. Ou, L. H. Vilela-Leo, D. C. Ralph, and\nR. A. Buhrman, Physical Review B 92, 064426 (2015).\n[23]J. Zhang, Y. Huai, L. Chen, and J. Zhang,\nJournal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures 21, 237 (2003) .\n[24]J.Slonczewski, Journal of Magnetism and Magnetic Materials 159, L1 (1996) .\n[25]A. Aharoni, Journal of Applied Physics 83, 3432 (1998) .\n[26]A. Vansteenkiste, J. Leliaert, M. Dvornik, M. Helsen,\nF. Garcia-Sanchez, and B. V. Waeyenberge,\nAIP Advances 4, 107133 (2014) .\n[27]A. Vansteenkiste and B. V. d. Wiele,\nJournal of Magnetism and Magnetic Materials 323, 2585 (2011) ." }, { "title": "0907.1800v1.Measurement_of_Conduction_Electron_Polarization_Via_the_Pairing_Resonance.pdf", "content": "Measurement of Conduction Electron Polarization Via the Pairing Resonance\nY.M. Xiong, P.W. Adams\nDepartment of Physics and Astronomy, Louisiana State University, Baton Rouge, Louisiana 70803, USA\nG. Catelani\nDepartment of Physics and Astronomy, Rutgers University, Piscataway, New Jersey 08854, USA\n(Dated: March 21, 2022)\nWe show that the pairing resonance in the Pauli-limited normal state of ultra-thin superconducting\nAl \flms provides a spin-resolved probe of conduction electron polarization in thin magnetic \flms. A\nsuperconductor-insulator-ferromagnet tunneling junction is used to measure the density of states in\nsupercritical parallel magnetic \felds that are well beyond the Clogston-Chandresekhar limit, thus\ngreatly extending the \feld range of tunneling density of states technique. The applicability and\nlimitations of using the pairing resonance as a spin probe are discussed.\nPACS numbers: 74.50.+r, 75.70.Ak, 85.75.-d,74.78.Db\nThe possibility of incorporating spin degrees of free-\ndom into electronic technologies has led to an explosion\nin research into mechanisms of spin polarization of con-\nduction currents in semiconducting and metallic systems\n[1, 2]. In addition to being technologically important,\nspin polarization in conducting systems remains a fun-\ndamentally interesting many-body problem. Clearly, an\naccurate determination of the conduction electron polar-\nizationP, particularly in thin-\flm magnetic structures,\nis crucial both from the point of view of \\spintronic\"\ndevice development and basic research. Unfortunately,\nhowever, there are few direct probes of P. The three\nmost successful techniques, listed in historical order, have\nbeen Zeeman-split superconducting tunneling density of\nstates spectroscopy (SCTDoS) [3, 4], spin-resolved pho-\ntoemission spectroscopy (SRPES) [5], and point contact\nAndreev re\rection (PCAR) [6, 7]. Each of these tech-\nniques has its own unique advantages and limitations.\nSCTDoS has outstanding resolution \u001810\u0016V, and is\ncompatible with thin-\flm geometries, but is limited to\na narrow range of magnetic \felds and is not easily im-\nplemented with bulk samples [4]. SRPES has a relatively\nlow resolution of\u001810 mV, and does not discriminate well\nbetween itinerant and localized bands. PCAR can be\nused on bulk samples and thick \flms, but it requires very\nlow impedance point contacts, and is also incompatible\nwith magnetic \felds [7, 8]. In this Letter we introduce\nan extension of SCTDoS technique that exploits the spin\nstructure of the Pauli-limited normal state (PLNS) pair-\ning resonance [9, 10] to measure electron polarization in\nmagnetic \felds well above the superconducting critical\n\feld. We show that the technique not only greatly ex-\ntends the range of \felds over which the polarization can\nbe measured, but it can be much less sensitive to \feld\nmisalignment than SCTDoS.\nTedrow and Meservey [4] pioneered the use of super-\nconducting spin-resolved tunneling to directly measure\nthe electron polarization in magnetic \flms. The tech-\nnique utilizes a planar junction geometry consisting ofa superconductor-oxide-ferromagnet (SC-Ox-FM) sand-\nwich in which the superconductor counter-electrode is\npurposely made very thin. The tunnel junction is then\ncooled below the superconducting transition temperature\nand a carefully aligned magnetic \feld is applied parallel\nto the junction plane. If the SC \flm thickness tis much\nless than the superconducting coherence length \u0018, then\nthe Meissner response to the applied \feld is suppressed,\nand the critical \feld of the SC counter-electrode is en-\ntirely Zeeman mediated [11]. If the spin-orbit scattering\nrate of the SC is small, then spin is a good quantum num-\nber, and the spin rotation symmetry of the SC can be ex-\nploited to provide a spin-resolved probe. In the Tedrow\nand Meservey technique the polarization is measured in\nthe superconducting phase by applying a subcritical par-\nallel magnetic \feld to the tunnel junction at low tem-\nperature,T\u001cTc. This induces a Zeeman splitting of\nthe SC quasiparticle DoS spectrum, and the BCS coher-\nence peaks get split into spin-up and spin-down bands\nby the parallel \feld [12]. As will be discussed below, the\nrelative heights of these peaks give a direct measure of\nthe electron polarization in the FM. Here we show that\npolarization can, in fact, be measured in \felds that are\nseveral times larger than the parallel critical of the SC\nby using the PLNS pairing resonance in appropriately\ndesigned tunnel junctions.\nThe SC-Ox-FM tunnel junctions were formed by \frst\ndepositing a thin Al counter-electrode via e-beam depo-\nsition of 99.999% Al stock onto \fre polished glass micro-\nscope slides held at 84 K. The depositions were made at\na rate of\u00180:1 nm/s in a typical vacuum with pressure\n<3\u000210\u00007Torr. After deposition, the counter-electrode\nwas exposed to the atmosphere for 10-20 minutes in or-\nder to allow a thin native oxide layer to form. Then\na 45 \u0017A thick FM \flm was deposited onto the counter-\nelectrode with the oxide serving as the tunneling barrier.\nIn this study the FM was either CNi 3or CCo 3, where\nthe e-beam depositions were made from arc-melted but-\ntons. The counter-electrode thicknesses were chosen soarXiv:0907.1800v1 [cond-mat.supr-con] 10 Jul 20092\nthat their in-plane sheet resistance was \u00181-2 k\n/sq. This\ncorresponded to thicknesses that were typically 22-24 \u0017A.\nThe low temperature parallel critical \felds of the counter-\nelectrodes were\u00186.5 T. The junction area was about\n1 mm\u00021 mm, while the junction resistance ranged from\n15-100 k\n depending on exposure time and other factors.\nOnly junctions with resistances much greater than that\nof the \flms were used. Measurements of resistance and\ntunneling were carried out on an Oxford dilution refriger-\nator using a standard ac four-probe technique. Magnetic\n\felds of up to 9 T were applied using a superconducting\nsolenoid. A mechanical rotator was employed to orient\nthe sample in situ with a precision of \u00180:1\u000e.\nIn the upper panel of Fig. 1 we plot the 70 mK tunnel-\ning conductance of a Al-AlO x-CNi 3tunnel junction in a\nsub-critical parallel magnetic \feld of 4 T. The Zeeman\nsplitting of the BCS DoS spectrum is clearly evident,\nwhere we have labeled the spin moment associated with\neach peak. Normally the respective spin peaks would be\nidentical on either side of the Fermi energy. The asym-\nmetry arises from the unequal spin populations in the\nferromagnetic CNi 3. This is expected since the magneti-\nzation properties of the CNi 3and CCo 3are very similar\nto their elemental counterparts [13]. If one assumes that\nspin is conserved in the tunneling processes, then tun-\nneling currents from the spin-up (down) bands in the Al\nwill only tunnel into corresponding spin-up (down) bands\nin the ferromagnet. Tedrow and Meservey exploited this\nconservation property to extract the polarization of a va-\nriety of transition metal FM \flms [4],\nP=\f\f\f\f\u000e1\u0000\u000e2\n\u000e1+\u000e2\f\f\f\f(1)\nwhere the peak height di\u000berences \u000e1;2are de\fned in\nFig. 1A and the corresponding polarization of the CNi 3\nis\u001818%.\nThe zero-\feld gap energy for the Al counter-electrode\nused in Fig. 1 was determined to be \u0001 o\u00180:49 meV by\n\fts to the superconducting DoS spectrum. Because the\nthickness of the Al counter-electrode is much less than\nthe superconducting coherence length \u0018\u0018150\u0017A, it un-\ndergoes a \frst-order parallel critical \feld transition at the\nClogston-Chandrasekhar critical \feld [14, 15]\nHCC\nc=\u0001op\n1 +G0\np\n2\u0016B(2)\nwhere\u0016Bis the Bohr magneton and G0is the anti-\nsymmetric Landau parameter [16]. In panel (B) of Fig. 1\nwe show the DoS spectrum at the critical \feld. This\nspectrum is in the coexistence region between the su-\nperconducting phase and the Pauli-limited normal state.\nThe new feature that appears at the transition is a mani-\nfestation of the pairing resonance (PR). Finally, in panel\n(C) we show a normal-state spectrum where the BCS\ncoherence peaks have been extinguished. The remain-\ning structure consists of a broad, symmetric, background\n0.80.850.90.9511.05\n-4-3-2-101234V (mV)H|| = 8.5 T(C)0.70.750.80.850.90.9511.05G(V)/G(4mV)H|| = 6.8 T(B)PR00.511.52\nH|| = 4.0 Tδ1δ2(A)FIG. 1: Evolution of the tunneling conductance of a Al-AlO x-\nCNi3 tunnel junction as the parallel critical \feld transition\nis crossed at 70 mK. A: superconducting phase showing an\nasymmetric Zeeman-split DoS spectrum. The arrows denote\nthe spin assignment of the coherence peaks. B: DoS spectrum\nat the parallel critical \feld transition in which the normal-\nstate pairing resonance (PR) coexists with superconducting\ncoherence peaks. C: Pauli-limited normal state in which only\nthe pairing resonance and the zero bias anomaly remain. Note\nthat the positive and negative resonances have di\u000berent mag-\nnitudes.3\nV*PLNSFM\nVH||EF-Ez/2VReverse BiasForward Bias+Ez/2PLNSPM\n(A)(B)\nFIG. 2: A: Schematic of the spin structure of the pairing\nresonance. B: DoS pro\fles of a tunnel junction comprised of\nan Al \flm in the Pauli-limited normal state (PLNS) on one\nside and a ferromagnetic \flm on the other. Note the depletion\nof states in the PLNS due to the pairing resonance and the\nzero bias anomaly.\nwith two small satellite resonances on either side of V =\n0. The background feature is often referred to as the zero\nbias anomaly and is a well documented electron-electron\ninteraction e\u000bect [17, 18]. In contrast, the satellite fea-\ntures represent incoherent Cooper pairing. The fact that\nthe resonance dips have unequal magnitude in Fig. 1 in-\ndicates that they are spin speci\fc. As we show below,\nthey can be used to extract FM polarization in \felds\nwell beyond the Clogston-Chandrasekahr critical \feld of\nEq. (2).\nFor the purposes of measuring polarization the most\nimportant property of the PR is its spin structure. In\nFig. 2 we present a graphic representation of the res-\nonance as it is observed in tunneling into a paramag-\nnetic (PM) metal \flm. Since the PM has no preferred\nspin direction, the resonance is symmetric about V = 0.\nAs is depicted in Fig. 2(A), when a su\u000ecient forward-\nbias voltage is reached, spin-down electrons in the PM\ncan tunnel across to form doubly occupied levels close to\nthe top of the spin-up band in the PLNS. These spin-\nsinglet states can then mix with the unoccupied states\nin the near vicinity to form an evanescent Cooper pair\n[9]. This e\u000bectively produces a small depletion of spin-\ndown quasiparticle states due to the fact that they have\nbeen consumed by the resonance. By particle-hole sym-\nmetry there is a similar depletion of spin-up states at\nthe reverse-bias voltage needed for spin-up electrons ly-\n-4 -3 -2 -1 0 1 2 3 4-0.10-0.08-0.06-0.04-0.020.00∆G/GV (mV)δ+δ‐\nFIG. 3: The red symbols are tunneling spectra taken at 70 mK\nin a parallel \feld of 6.7 T, where the zero bias anomaly back-\nground has been subtracted o\u000b. The orange arrows denote the\nspin assignment of the occupied and unoccupied resonances.\nThe solid black line represents a best \ft to the resonance\ncurve. The dashed blue line is the predicted resonance pro\fle\nat 18 T, extrapolating from the \ftting parameters obtained\nat 6.7 T.\ning just above the PLNS doubly occupied sites to tunnel\nover to the top of the spin-up band in the PM. The pre-\ncise energy of these resonances is \feld dependent,\neV\u0003=1\n2\u0012\nEz+q\nE2z\u0000\u00012\n0\u0013\n; (3)\nwhereEz= (2\u0016BH)=(1 +G0) is the Zeeman energy\nrenormalized by G0.\nThe zero bias anomaly background in Fig. 1(C) is\nindependent of magnetic \feld and varies as ln Vfor\nV&kBT=e. It can easily be subtracted from the data in\norder to isolate the resonances as shown in Fig. 3. The\nred dots are data taken on a Al-AlO x-CCo 3tunnel junc-\ntion at 70 mK in a 6.7 T parallel magnetic \feld. The ar-\nrows depicted the spin assignments of the resonances and\n\u000e\u0006refer to their respective amplitudes. The solid black\nline is a best \ft to the resonance pro\fle using a procedure\nand formalism described elsewhere [19]. As was the case\nfor the superconducting phase, we only need the relative\nmagnitudes of the positive and negative bias resonances\nin order to determine the polarization. If we take the\nPLNS density of states to be Ns, then well away from the\nPR spin rotation symmetry requires Ns\n\"=Ns\n#=Ns=2.\nOn resonance, however, there will be a depletion of one\nspin component at positive bias and the other at negative\nbias,\nNs\n+;\u0000=Ns\n#;\"+Ns\n\";#(1\u0000\u000f) (4)\nwhere 0\u0014\u000f\u00141 represents the strength of the resonance.\nIn general\u000fdepends upon a number of factors, including\nmagnetic \feld, temperature, spin-orbit scattering rate,4\n0102030405060\n0246810CCo3CNi3P (%) \nH (T)\nFIG. 4: Electron polarization of 45 \u0017A CNi 3and CCo 3\flms as\na function of magnetic \feld. The dashed line represents the\napproximate parallel critical \feld of the Al counter-electrodes\nused in these measurements.\nand the dimensionless normal-state transport conduc-\ntance,g[9, 19]. Spin conservation forbids intra-spin-band\ntunneling, so if we assume that the ferromagnet has a ma-\njority spin density Nf\n\"and a minority spin density Nf\n#,\nthen the tunneling conductance is simply proportional to\nthe product of respective spin-speci\fc DoS on either side\nof the tunnel junction as is depicted in Fig. 2(B). The\nmagnitudes of the positive- and negative-bias resonance\nfeatures are\n\u000e+;\u0000= [Ns\n#;\"Nf\n#;\"+Ns\n\";#Nf\n\";#(1\u0000\u000f)]\n\u0000[Ns\n#;\"Nf\n#;\"+Ns\n\";#Nf\n\";#]\n=\u0000\u000fNsNf\n\";#\n2:(5)\nFrom this the polarization follows,\nP=\f\f\f\f\fNf\n\"\u0000Nf\n#\nNf\n\"+Nf\n#\f\f\f\f\f=\f\f\f\f\u000e+\u0000\u000e\u0000\n\u000e++\u000e\u0000\f\f\f\f: (6)\nThe data in Fig. 3 give a polarization of 39.5% which\nis comparable to values reported for pure Co by the early\nwork of Tedrow and Meservey [4] and later measurements\nusing PCAR [7]. We note that Eq. (6) is independent of\nboth the PR's strength \u000fand its width. Thus within the\nlimits of signal-to-noise constraints, one should be able\nto obtain polarization measurements at quite high \felds.\nThe dashed line in Fig. 3 represents a prediction for the\nresonance curve at 18 T obtained by extrapolating thenecessary parameters from \ftting the data at 6.7 T. Al-\nthough the PR is signi\fcantly attenuated and broadened\nby the high \feld, it is still well within the noise level\nof the data. Since the strength of the resonance grows\nas the conductance gis reduced, one can use slightly\nthinner counter-electrodes to increase the visibility of the\nresonance. Previous studies have shown that the PR in\ncounter-electrodes with g\u00146 remains well de\fned in\nperpendicular \felds of a few Tesla, thus eliminating the\nneed for precise parallel alignment [20].\nIn Fig. 4 we present electron polarization as a function\nof magnetic \feld for CNi 3and CCo 3\flms. The verti-\ncal dashed line represents the approximate critical \feld\nof the counter-electrodes used in this study. The data\nshow good agreement between polarization values mea-\nsured just below Hcjjand normal-state values measured\njust above Hcjj. The polarization of CNi 3\flm is inde-\npendent of \feld, as would be expected considering the\nfact that the Zeeman energies associated with \felds of\na few Tesla are less than 0.1 meV, which is much less\nthan exchange energies associated with the magnetiza-\ntion. Interestingly, though, the polarization of the CCo 3\n\flm exhibits a signi\fcant decrease in \felds above 3-4 T.\nThis trend can be seen in both the superconducting phase\nmeasurements and the PR measurements. The origin of\nthis \feld dependence is uncertain.\nIn summary, we have shown that conduction spin po-\nlarization can be determined from the relative amplitudes\nof the occupied and unoccupied pairing resonance fea-\ntures. The technique can be used in \felds that are sev-\neral times higher than the Clogston-Chandrasehkar crit-\nical \feld, thus allowing polarization measurements to be\nmade over a very wide range of magnetic \felds. Prelim-\ninary polarization measurements in CCo 3\flms show a\nstrong suppression of the polarization in \felds above a\nfew Tesla.\nWe gratefully acknowledge enlightening discussions\nwith Ilya Vekhter and Dan Sheehy. This work was\nsupported by the DOE under Grant No. DE-FG02-\n07ER46420.\n[1] S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M.\nDaughton, S. von Molnr, M. L. Roukes, A. Y. Chtchelka-\nnova, and D. M. Treger, Science 294, 1488 (2001).\n[2] G. Prinz, Phys. Today 48, 58 (1995).\n[3] P.M. Tedrow and R. Meservey, Phys. Rev. Lett. 26, 192\n(1971).\n[4] P.M. Tedrow and R. Meservey, Phys. Rep. 238, 173\n(1994).\n[5] P.D. Johnson, Rep. Prog. Phys. 60, 1217 (1997).\n[6] S.K. Upadhyay, A. Palanisami, R.N. Louie, and R.A.\nBuhrman, Phys. Rev. Lett. 81, 3247 (1998).\n[7] J. Soulen, Jr., J. M. Byers, M. S. Osofsky, B. Nadgorny,\nT. Ambrose, S. F. Cheng, P. R. Broussard, C. T. Tanaka,\nJ. Nowak, J. S. Moodera, A. Barry, and J. M. D. Coey,5\nScience 282, 85 (1998).\n[8] P. Chalsani, S.K. Upadhyay, O. Ozatay, and R.A.\nBuhrman, Phys. Rev. B 75, 094417 (2007).\n[9] I. L. Aleiner and B. L. Altshuler, Phys. Rev. Lett. 79,\n4242 (1997).\n[10] V. Y. Butko, P. W. Adams, and I. L. Aleiner, Phys. Rev.\nLett. 82, 4284 (1999); H. Y. Kee, I. L. Aleiner, and B. L.\nAltshuler, Phys. Rev. B 58, 5757 (1998).\n[11] P. Fulde, Adv. Phys. 22, 667 (1973).\n[12] R. Meservey, P.M. Tedrow, and P. Fulde, Phys. Rev. Lett.\n25, 1270 (1970).\n[13] D.P. Young, A.B. Karki, P.W. Adams, J.N. Ngunjiri, J.C.\nGarno, H. Zhu, B. Wei, D. Moldovan, J. App. Phys. 103,\n053503 (2008).[14] A. M. Clogston, Phys. Rev. Lett. 9, 266 (1962).\n[15] B. S. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962).\n[16] G. Catelani, X. S. Wu, and P. W. Adams, Phys. Rev. B\n78, 104515 (2008).\n[17] B.L. Altshuler, A.G. Aronov, M.E. Gershenson, and\nYu.V. Sharvin, Sov. Sci. Rev. A. Phys. Vol. 9, 223 (1987).\n[18] Y. Imry and Z. Ovadyahu, Phys. Rev. Lett. 49, 841\n(1982).\n[19] G. Catelani, Y. M. Xiong, X. S. Wu, P. W. Adams, e-\nprint arXiv:0905.2414.\n[20] X.S. Wu, P.W. Adams, and G. Catelani, Phys. Rev. Lett.\n95, 167001 (2005)." }, { "title": "1912.02337v2.Spin_pumping_into_a_spin_glass_material.pdf", "content": "Spin pumping into a spin glass material\nYusei Fujimoto,1Masanori Ichioka,2, 1and Hiroto Adachi2, 1\n1Department of Physics, Okayama University, Okayama 700-8530, Japan\n2Research Institute for Interdisciplinary Science, Okayama University, Okayama 700-8530, Japan\n(Dated: March 24, 2022)\nSpin pumping is a recently established means for generating a pure spin current, whereby spins\nare pumped from a magnet into the adjacent target material under the ferromagnetic resonance\ncondition. We theoretically investigate the spin pumping from an insulating ferromagnet into spin\nglass materials. Combining a dynamic theory of spin glasses with the linear-response formulation\nof the spin pumping, we calculate temperature dependence of the spin pumping near the spin glass\ntransition. The analysis predicts that a characteristic peak appears in the spin pumping signal,\nre\recting that the spin \ructuations slow down upon the onset of spin freezing.\nI. INTRODUCTION\nSpin current is a \row of spin angular momentum [1].\nOver the last two decades, great progress has been made\nin generating, manipulating, and detecting the spin cur-\nrent [2, 3]. With regard to the spin current generation,\nas nicely reviewed in Ref. [4] the spin pumping is now\nestablished as a charge-free and versatile means [5{8]. In\nthis method a pure spin current, which is unaccompa-\nnied by a charge current, is pumped from a ferromagnet\ninto the adjacent spin sink material by a stimulus of mi-\ncrowaves satisfying the ferromagnetic resonance (FMR)\ncondition. Thanks to the advent of the spin pumping\ntechnique, the spin current physics has so far been inves-\ntigated in a variety of spin sink materials, ranging from\nnonmagnetic metals [9{13], semiconductors [14{16], mag-\nnetic metals [17{19], insulators [20], to more exotic sys-\ntems such as graphene [21{23], transition metal dichalco-\ngenides [24], organic materials [25, 26], and strongly spin-\norbit coupled materials [27, 28].\nRecently, the playground of the spin current physics\nhas been extended to disordered magnets or the so-called\nspin glass (SG) materials [29, 30]. The SGs are charac-\nterized by a freezing of random spins [31], and its nature\nhas long been studied both experimentally and theoreti-\ncally [32]. However, despite its long history of research,\nthe interplay of spin current and the SG ordering has\nnot yet been well examined. Thus, it is quite natural to\nask what happens if we inject a pure spin current into a\nSG material by the spin pumping. Experimentally, the\nspin pumping into a SG material was reported in 2011\nusing a Ag 90Mn10/Ni81Fe19bilayer [33]. To the best of\nour knowledge, however, no theoretical work on the spin\npumping into SG materials can be found in the literature.\nTherefore, developing a theory of spin pumping into the\nSG material is highly desirable.\nIn this paper, we theoretically investigate the spin\npumping into SG materials. Although a metallic magnet\nNi81Fe19was used as the spin injecting magnet in the\nprevious experiment [33], we consider here an insulat-\ning ferromagnet such as yttrium iron garnet for the spin\ninjector, since it makes the spin pumping signal more vis-\nible. Our strategy to calculate the spin pumping into SGmaterials is as follows. First, we use a linear-response\napproach to the spin pumping [34, 35]. The notion de-\nrived from the linear response approach, that the spin\npumping is intimately related to the dynamic spin sus-\nceptibility of the spin sink layer, has successfully been\napplied to the spin pumping into a ferromagnet [36], an-\ntiferromagnets [37, 38], and recently it was also applied\nto the spin pumping into superconductors [39{41]. Thus,\nwe relate the spin pumping signal to the dynamic spin\nsusceptibility of the SG layer. Next, we calculate the sus-\nceptibility of the SG layer by employing a dynamic theory\nof SGs [42, 43]. Not only that this theory is known to\nbe an alternative formulation of the static replica the-\nory [44{47], but also that the dynamic theory is more\nsuitable to discuss the dynamic quantity such as the dy-\nnamic spin susceptibility [48].\nIn the literature, the dynamic spin susceptibility near\nthe SG transition was calculated [48], but the result\nwas limited to the Ising case and to an extremely low-\nfrequency regime less than 10 KHz, which is out of the\nFMR condition. In the present paper, we extend the sus-\nceptibility calculation of Ref. [48] to the Heisenberg case\nand GHz frequency regime that is relevant to spin pump-\ning experiments, and calculate temperature dependence\nof the spin pumping into SG materials. With this, we\nshow that a characteristic peak structure appears near\nthe SG transition, which is a consequence of the slowing\ndown of spin \ructuations that is concomitant with the\nspin freezing of the system.\nThe plan of this paper is as follows. In the next section,\nwe introduce our microscopic model, and relate the spin\npumping with the dynamic spin susceptibility of impu-\nrity spins. In Sec. III, on the basis of the dynamic theory\nof SGs, we explain how to calculate the dynamic spin sus-\nceptibility of impurity spins. In Sec. IV, the spin pump-\ning signal into a SG material is calculated as a function\nof temperature. Finally, in Sec. V we discuss and sum-\nmarize our results.arXiv:1912.02337v2 [cond-mat.mes-hall] 13 May 20202\nFIG. 1. Schematics of the system considered in this paper,\nwhere the bilayer is composed of a SG material and a fer-\nromagnetic insulator (FI). Here, \u001bandSare, respectively,\nthe conduction-electron spin and the impurity spin in the SG\nlayer, and \nis the localized spin in the FI layer. A spin cur-\nrent with a helicity opposite to \n\rows from the FI layer to\nthe SG layer.\nII. MODEL\nWe consider a bilayer composed of a ferromagnetic in-\nsulator (FI) and a SG material, as shown in Fig. 1. More\nconcretely, we may think of yttrium iron garnet (YIG)\nfor the FI layer and Mn-doped Cu (Cu:Mn) for the SG\nlayer. We assume that a static magnetic \feld H0=H0^ z\nis applied to the FI/SG bilayer in the lateral direction,\nand that the anisotropy \feld is much smaller than H0\nsuch that it can be discarded.\nWe start from the following Hamiltonian:\nH=HSG+HFI\u0000SG; (1)\nwhere the \frst term,\nHSG=X\np\u0018pcy\npcp+JeSX\nra\u001b(ra)\u0001S(ra); (2)\ndescribes the SG layer [49, 50]. Here, the \frst term on the\nright-hand side describes the conduction electron kinetic\nenergy, and the second term the coupling between the\nconduction electron spin and magnetic impurity at an\nimpurity position ra, whereJeSis thesd-type exchange\ncoupling. Here, cy\np= (cy\np;\";cy\np;#) is the electron creation\noperator for spin projection \"and#,Sis an impurity\nspin,\u001b(r) =cy(r)^\u001bc(r) is the spin density operator with\n^\u001bbeing the Pauli matrices, and c(r) =N\u00001=2\nSGP\npcpeip\u0001r\nwithNSGbeing the number of lattice sites at the SG\nlayer.\nThe second term of Eq. (1),\nHFI\u0000SG=JintX\nrint\u001b(rint)\u0001\n(rint); (3)\ndescribes the interaction between the FI and SG layers.\nHere,Jintis the interfacial sdcoupling between the con-\nduction electron spins in the SG and the localized spins\nin the FI, where rintis a position at the FI/SG interface.\nSG\nFIFIG. 2. Diagrammatic representation of the magnon self-\nenergy giving the spin pumping signal. Here, \u001f(\u001b)is\nthe dynamic spin susceptibility of conduction-electron spins,\nwhereas\u001f(S)is that of impurity spins.\nIn order to investigate the spin pumping in the present\nsystem, we use the linear-response formulation of the spin\npumping [34, 35]. We consider the situation where an\nexternal microwave with the angular frequency !acis ap-\nplied to the FI/SG bilayer that drives the FMR of the\nFI side. The linear-response formulation uses the follow-\ning magnon language. In the absence of the adjacent\nSG layer, the uniform-mode (Kittel mode) magnon has\nan intrinsic damping rate \u000b0!ac, where\u000b0is the intrin-\nsic Gilbert damping constant. In the presence of the SG\nlayer, since the spin-relaxation rate due to the SG layer is\nadditive and hence an additional spin dissipation channel\nopens, there arises an additional magnon damping rate.\nTherefore, the total Gilbert damping constant \u000bfor the\nbilayer is given by\n\u000b=\u000b0+\u000e\u000b; (4)\nwhere\u000e\u000bis the additional Gilbert damping constant.\nThe relationship between this additional Gilbert damp-\ning constant and the spin current Ispumped into the SG\nlayer withz-axis polarization is given by [34]\nIs=\u000e\u000b\r~\nMsV!ac(\rhac)2\n(\rH0\u0000!ac)2+ (\u000b0!ac)2; (5)\nwhere\ris the gyromagnetic ratio, Msthe saturation\nmagnetization, Vthe volume of the magnet, and hacis\nthe amplitude of the external microwave.\nAccording to the linear-response formulation [34, 35],\nthe additional magnon damping rate can be calculated\nfrom the corresponding magnon self-energy \u0006( !) whose\nprocess involves the spin transfer across the interface\n(Fig. 2). In the present situation, up to the lowest or-\nder with respect to Jint, the self-energy is given by\n\u0006(!) =\u0000J2\nintNint\n~2NSGNFIX\nq\u001f(\u001b)\nq(!)JeS\u001f(S)\nq(!)JeS\u001f(\u001b)\nq(!);\n(6)\nwhereNintis the number of the localized spins \nat the\nFI/SG interface, and NFIis the number of lattice sites3\nin the FI layer. In the above equation, \u001f(\u001b)\nq(!) is the\nFourier transform of the retarded susceptibility of the\nconduction-electron spin \u001b, i.e.,\u001f(\u001b)\nq(t\u0000t0) = i\u0002(t\u0000\nt0)h[\u001b\u0000\nq(t);\u001b+\n\u0000q(t0)]i, where \u0002(t) is the step function and\nwe de\fned O\u0006=Ox\u0006iOyfor a vector operator O. By\ncontrast,\u001f(S)\nq(!) is the Fourier transform of the retarded\nsusceptibility of the impurity spin S, i.e.,\u001f(S)\nq(t\u0000t0) =\ni\u0002(t\u0000t0)h[S\u0000\nq(t);S+\n\u0000q(t0)]i,\nUsing the relation \u000e\u000b=\u0000!\u00001\nacIm\u0006(!ac) [34], the addi-\ntional Gilbert damping constant \u000e\u000bis expressed as\n\u000e\u000b\u0019J2\nintNint\n~2NFI\u0010\n\u001f(\u001b)\n0JeS\u001121\n!acIm\u001f(S)\nloc(!ac); (7)\nwhere we introduced a shorthand notation \u001f(\u001b)\n0=\n\u001f(\u001b)\nq=0(!= 0), and\u001f(S)\nloc(!) =N\u00001\nSGP\nq\u001f(S)\nq(!) is the local\nsusceptibility of impurity spins. In obtaining the above\nresult, we made use of the fact that \u001f(S)\nloc(!) is de\fned in\na smallqregionq.2\u0019=bwherebis the average distance\nof two magnetic impurities. In this small qregion, the\nconduction-electron spin susceptibility is approximated\nby the uniform and static component \u001f(\u001b)\n0, where this\nquantity is pure real.\nEquation (7) means that the additional Gilbert damp-\ning constant \u000e\u000bdue to the spin pumping is proportional\nto the imaginary part of the dynamic spin susceptibil-\nity\u001f(S)\nloc(!). In this expression, the strongest temper-\nature dependence upon the SG transition results from\nthe imaginary part of the dynamic spin susceptibility,\nIm[\u001f(S)\nloc(!)]. This means that, as long as the temperature\ndependence is concerned, the part other than Im[ \u001f(S)\nloc(!)]\ncan be regarded as being temperature independent, and\nthe temperature dependence is dominated by that of\nIm[\u001f(S)\nloc(!)]. We adopt this approximation in the nu-\nmerical calculation in Sec. IV.\nThe quantity \u001f(S)\nloc(!) is a correlation function between\ntwo impurity spins, and hence it can be evaluated using\nour knowledge on SGs. In the next section, we evaluate\n\u001f(S)\nloc(!) using a dynamic theory of SGs [42, 43].\nIII. DYNAMIC SPIN SUSCEPTIBILITY OF\nIMPURITY SPINS\nIn this section, by employing a dynamic theory of\nSGs [42, 43], we sketch our procedure for calculating\nthe dynamic spin susceptibility \u001f(S)\nloc(!) appearing in the\nspin pumping signal [Eq. (7)]. We emphasize that we\nuse the dynamic model of the Heisenberg spin glasses\ndeveloped by Sompolinsky and Zippelius [43], which is\nconstructed on top of a similar dynamic theory of Ising\nspin glasses [42].\nWe \frst integrate out the conduction-electron degrees\nof freedom in the Hamiltonian for the SG layer HSG.Then,HSGis transformed into the following form [49, 50]:\nHSG=1\n2X\ni6=jJijS(ri)\u0001S(rj); (8)\nwhereJijis nominally the RKKY interaction of the form\nJij=JeScos(2kFrij)=r3\nijwithrij=jri\u0000rjj. However,\nfollowing the standard approach to the SG problem [32],\nwe regards Jijas Gaussian random variables with zero\nmean and variance [ J2\nij]av=J2=NS, where [\u0001\u0001\u0001]avmeans\nthe random average over the distribution of Jij, andNS\nis the number of impurity spins.\nHamiltonianHSGin Eq. (8) is the same as the vec-\ntor spin version [51] of the Sherrington-Kirkpatrick (SK)\nmodel [45], so that we employ the established dynamical\napproach. Following Sompolinsky and Zippelius [43], we\n\frst replaceHSGwith its soft-spin version:\n\feHSG=1\n2X\n\u000bX\nij(r0\u000eij\u0000\fJij)S\u000b\niS\u000b\nj\u0000\fX\nih\u000b\ni\u0001S\u000b\ni;(9)\nwhere\f= 1=kBTis the inverse temperature, and each\ncomponent of the soft spin varies \u000011 ps). There the magnetization dynamics for\nthe left and right polarized pulses become symmetric be-\ncause for the right polarized \feld, the helicity of the light\nand the magnetization direction are aligned for the whole\ndynamics and for the left polarized case the helicity an\nmagnetization are aligned in the same \u0000zdirection.\nFinally, we would like to investigate the connection to\nperturbation theory with respect to the external \feld,\nwhich is often regarded as the sole contribution of the\ncoherent optical \feld. We can achieve this by computing\nthe magnetization dynamics for a \fxed band structure.\nIn this case, we do not update the eigenenergies and\neigenstates of hamiltonian (1), i.e., we keep the Stoner\nmean-\feld spliting \fxed, so that the change in magneti-\ntime [ps]spin polarization\n0 0.1 0.2 0.3 0.4 0.50.330.340.350.360.37FIG. 3. Magnetization (spin polarization) dynamics com-\nputed without carrier redistribution and for a \fxed band\nstructure with a left-polarized (solid line) and a right-\npolarized (dashed line) optical \feld. The CW optical \feld\nis closer to resonance with a detuning of 5 meV. This setup\nis designed to approximate the change of the density matrix\ndue to the coherent optical \feld and leads to a reduction in\nmagnetization for both circular polarizations of the optical\n\feld.\nzation comes from the dynamics of the reduced density\nmatrix under the in\ruence of the optical \feld. Such a\ncalculation for right and left-circularly polarized optical\n\feld is shown in Fig. 3. For a detuning of 200 meV the\ne\u000bect on the magnetization is vanishingly small, so that,\nin addition to keeping the band structure \fxed, we now\nuse a \feld closer to the resonance between magnetic and\nnon-essential bands, namely a detuning of 5 meV. The\nother parameters are kept the same as in Fig. 2. In this\nscenario, the magnetization approaches di\u000berent steady\nstates with a reduced magnetization (compared to equi-\nlibrium) for both right and left-polarized optical \felds.\nAs the density matrix includes the in\ruence of the co-\nherent \feld to all orders of the \feld, this result should\ncorrespond to a mechanism for the inverse Faraday e\u000bect\nanalyzed by Oppeneer and coworkers in the framework\nof 2nd-order perturbation theory for close-to-resonance\n\felds. Indeed, Berritta et al.36have also found that the\nin\ruence of optical \felds with opposite circular polariza-\ntion may lead to a reduced steady-state magnetization\nwith respect to the equilibrium magnetization. This is\nsomewhat contrary to what one expects from older the-\nories of the inverse Faraday e\u000bect which yielded exactly\nopposite e\u000bective magnetic \felds for opposite circular po-\nlarizations of the optical \feld. This \\antisymmetry\" be-\ntween the magnetic \feld and the circular optical polar-\nization is not present because of the \fnite equilibrium\nmagnetization, which breaks the symmetry between the\ndynamics induced by right and left-circularly polarized\n\felds.37In our case, we \fnd that the asymmetry is pro-\nnounced for the close-to-resonance case in Fig. 3 because\nthere the optical \feld \\sees\" the ferromagnetic splitting.\nIn the o\u000b-resonant case in Fig. 2(c), the ferromagnetic\nsplitting should play a smaller role, and indeed the mag-\nnetization dyamics are closer to realizing antisymmetry,\nas the magnetization is changed in opposite directions for5\nopposite polarizations.\nConclusion. We introduced a microscopic dynamical\nmodel to study the all-optical magnetization switching\nprocess in a simple ferromagnetic band structure, includ-\ning spin-orbit coupling and incoherent carrier redistribu-\ntion/scattering processes. For o\u000b-resonant excitation we\nfound that the switching process is a combination of de-\nmagnetization and the in\ruence of the o\u000b-resonant \feld\nin the form of a dynamical Stark e\u000bect. The main an-\ngular momentum change is supplied by the lattice via\nan Elliott-Yafet like magnetization change, which results\nfrom the combination of spin-orbit coupling and scatter-\ning processes. Even for a continuous-wave excitation, we\nfound that the \feld acts directly on the magnetization\nvia an o\u000b-resonant Raman-like processes, which is closelyrelated to the dynamical Stark e\u000bect. This Raman-like\nprocess leads to a decrease/increase of the electronic spin\npolarization, i.e., the magnetization, as one would expect\nfrom the angular momentum supplied by the left/right\ncircularly optical \feld. However, both circular polariza-\ntions increase the energy of the electrons in the spin-split\nbands during the duration of the optical \feld, and thus\ncontribute to a demagnetization e\u000bect largely indepen-\ndent of the polarization. For optical \felds close to res-\nonance we \fnd magnetization changes that are not an-\ntisymmetric with respect to the optical polarization, in\nagreement with recent perturbation theory calculations.\nSvenja Vollmar received a fellowship through the Ex-\ncellence Initiative (DFG/GSC 266). We acknowledge\nsupport from DFG by the SFB/TRR Spin+X.\n\u0003Graduate School of Excellence Materials Science in Mainz,\nGottlieb-Daimler-Strasse 47, 67663 Kaiserslautern, Ger-\nmany\nyhcsch@physik.uni-kl.de\n1E. Beaurepaire, J.-C. Merle, A. Daunois, and J.-Y. Bigot,\nPhysical Review Letters 76, 4250 (1996).\n2I. Radu, K. Vahaplar, C. Stamm, T. Kachel, N. Pontius,\nH. A. D urr, T. A. Ostler, J. Barker, R. F. Evans, R. W.\nChantrell, et al. , Nature 472, 205 (2011).\n3J. H. Mentink, J. Hellsvik, D. V. Afanasiev, B. A. Ivanov,\nA. Kirilyuk, A. V. Kimel, O. Eriksson, M. I. Katsnelson,\nand T. Rasing, Phys. Rev. Lett. 108, 057202 (2012).\n4T. A. Ostler, J. Barker, R. F. Evans, R. W. Chantrell,\nU. Atxitia, O. Chubykalo-Fesenko, S. El Moussaoui,\nL. Le Guyader, E. Mengotti, L. J. Heyderman, et al. , Nat.\nCommun. 3, 666 (2012).\n5S. Wienholdt, D. Hinzke, K. Carva, P. M. Oppeneer, and\nU. Nowak, Physical Review B 88, 020406 (2013).\n6A. Baral and H. C. Schneider, Phys. Rev. B 91, 100402(R)\n(2015).\n7S. Mangin, M. Gottwald, C.-H. Lambert, D. Steil, V. Uhl\u0013 \u0010\u0014 r,\nL. Pang, M. Hehn, S. Alebrand, M. Cinchetti, G. Mali-\nnowski, Y. Fainman, M. Aeschlimann, and E. E. Fullerton,\nNat. Mater. 13, 286 (2014).\n8R. John, M. Berritta, D. Hinzke, C. M uller, T. San-\ntos, H. Ulrichs, P. Nieves, J. Walowski, R. Mondal,\nO. Chubykalo-Fesenko, J. McCord, P. M. Oppeneer,\nU. Nowak, and M. M unzenberg, Sci. Rep. 7, 4114 (2017).\n9A. V. Kimel, A. Kirilyuk, P. A. Usachev, R. V. Pisarev,\nA. M. Balbashov, and T. Rasing, Nature 435, 655 (2005).\n10A. Kirilyuk, A. V. Kimel, and T. Rasing, Rev. Mod. Phys.\n82, 2731 (2010).\n11S. Alebrand, A. Hassdenteufel, D. Steil, M. Cinchetti, and\nM. Aeschlimann, Phys. Rev. B 85, 092401 (2012).\n12A. Hassdenteufel, J. Schmidt, C. Schubert, B. Hebler,\nM. Helm, M. Albrecht, and R. Bratschitsch, Phys. Rev. B\nCondens. Matter 91, 104431 (2015).\n13L. P. Pitaevskii, Sov. Phys. JETP 12, 1008 (1961).\n14P. Pershan, J. Van der Ziel, and L. Malmstrom, Physical\nReview (1966).\n15R. Hertel, Journal of magnetism and magnetic materials\n303, L1 (2006).\n16M. I. Kurkin, N. B. Bakulina, and R. V. Pisarev, Phys.Rev. B Condens. Matter 78, 134430 (2008).\n17S. R. Woodford, Phys. Rev. B Condens. Matter 79, 212412\n(2009).\n18K. Taguchi and G. Tatara, Phys. Rev. B Condens. Matter\n84, 174433 (2011).\n19R. Hertel and M. F ahnle, Phys. Rev. B Condens. Matter\n91, 020411 (2015).\n20V. M. Edelstein, Phys. Rev. Lett. 80, 5766 (1998).\n21A. Qaiumzadeh and M. Titov, Phys. Rev. B 94, 014425\n(2016).\n22N. I. Zheludev, M. A. Brummell, R. T. Harley, A. Mali-\nnowski, S. V. Popov, D. E. Ashenford, and B. Lunn, Solid\nState Commun. 89, 823 (1994).\n23R. V. Mikhaylovskiy, E. Hendry, and V. V. Kruglyak,\nPhys. Rev. B Condens. Matter 86, 100405 (2012).\n24M. Battiato, G. Barbalinardo, and P. M. Oppeneer, Phys.\nRev. B Condens. Matter 89, 014413 (2014).\n25M. Berritta, R. Mondal, K. Carva, and P. M. Oppeneer,\nPhys. Rev. Lett. 117, 137203 (2016).\n26D. Popova, A. Bringer, and S. Bl ugel, Phys. Rev. B 85,\n094419 (2012).\n27A. Qaiumzadeh, G. E. W. Bauer, and A. Brataas, Phys.\nRev. B Condens. Matter 88, 064416 (2013).\n28H. Haug and S. W. Koch, Quantum theory of the optical\nand electronic properties of semiconductors (World Scien-\nti\fc Publishing Co Inc, 2009).\n29C. Scholl, S. Vollmar, and H. C. Schneider, \\Light-induced\nmagnetization dynamics in a ferroamgnetic rashba model,\"\n(2018), unpublished.\n30D. Popova, A. Bringer, and S. Bl ugel, Phys. Rev. B Con-\ndens. Matter 84, 214421 (2011).\n31B. J. Sussman, American Journal of Physics 79, 477\n(2011), https://doi.org/10.1119/1.3553018.\n32M. Krau\u0019, T. Roth, S. Alebrand, D. Steil, M. Cinchetti,\nM. Aeschlimann, and H. C. Schneider, Phys. Rev. B 80,\n180407 (2009).\n33B. Y. M uller, A. Baral, S. Vollmar, M. Cinchetti,\nM. Aeschlimann, H. C. Schneider, and B. Rethfeld, Phys.\nRev. B 111, 167204 (2013).\n34K. Leckron, S. Vollmar, and H. C. Schneider, Phys. Rev.\nB96, 140408 (2017).\n35M. S. El Hadri, P. Pirro, C. H. Lambert, S. Petit-Watelot,\nY. Quessab, M. Hehn, F. Montaigne, G. Malinowski, and6\nS. Mangin, Phys. Rev. B: Condens. Matter Mater. Phys.\n94, 064412 (2016).\n36M. Berritta, R. Mondal, K. Carva, and P. M. Oppeneer,\nPhys. Rev. Lett. 117, 137203 (2016).37M. B. Agranat, S. I. Ashikov, A. B. Granovskii, and G. I.\nRukman, Zh. Eksp. Teor. Fiz 86, 1376 (1986)." }, { "title": "1803.05038v1.Ultrafast_light_switching_of_ferromagnetism_in_EuSe.pdf", "content": "arXiv:1803.05038v1 [cond-mat.mtrl-sci] 13 Mar 2018DRAFTUltrafast light switching of ferromagnetism in EuSe\nA. B. Henriques,1X. Gratens,1P. A. Usachev,1∗V. A. Chitta,1and G. Springholz2\n1Instituto de F´ ısica, Universidade de S˜ ao Paulo, 05508-09 0 S˜ ao Paulo, Brazil\n2Institut f¨ ur Halbleiter und Festk¨ orperphysik, Johannes Kepler Universit¨ at Linz, 4040 Linz, Austria\nWe demonstrate that light resonant with the bandgap forces t he antiferromagnetic semiconducor\nEuSe to enter ferromagnetic alignment in the picosecond tim e scale. A photon generates an electron-\nhole pair, whose electron forms a supergiant spin polaron of magnetic moment of nearly 6,000 Bohr\nmagnetons. By increasing the light intensity, the whole of t he sample can be fully magnetized.\nThe key to the novel large photoinduced magnetization mecha nism is the huge enhancement of the\nmagnetic susceptibility when both antiferromagnetic and f erromagnetic interactions are present in\nthe material, and are of nearly equal magnitude, as is the cas e in EuSe.\nPACS numbers:\nThe ultrafast control of the magnetic state of matter\nis a topic of vast current interest for the development of\napplications and for advancing the knowledge in the field\nof light-matter interaction [1–3]. Here we report on the\ndiscovery of an ultrafast mechanism of light switching\nof antiferromagnetic EuSe into the ferromagnetic phase,\nwhichisefficientintheproximityofitsN´ eeltemperature.\nThe EuSe samples were grown by molecular beam epi-\ntaxy (MBE) onto (111) BaF 2substrates. Because of the\nalmost perfect lattice constant matching ( a= 6.191˚A\nanda= 6.196˚AforEuSeandBaF 2, respectively), nearly\nunstrained bulklike EuSe reference layers with µm thick-\nness were obtained directly by growth on BaF 2. The\ntime-resolved photoinduced Faraday was measured using\natwo-colorpump-probetechnique using 2ps light pulses,\nasillustrated in Fig. 1. The Faradayrotationangleofthe\nprobe light pulse, ∆ θF, induced by the pump pulse, was\nmeasured using lock-in techniques with a resolution bet-\nter than 10−7radians. The image of the excitation spot\non the sample had a diameter of 150 µm, about twice the\ndiameter of the probe spot. For continuous wave (CW)\nmeasurements the same setup was used, except that the\npump source was a doubled Nd:Yag laser or a Xe lamp\ncoupled to a monochromator, and the probe source was a\nsemiconductor laser of energy below the EuSe bandgap.\nAll measurements were performed using an optical cryo-\nstat containing a superconducting coil for magnetic fields\nup to 8 Teslas, applied in the Faraday geometry. Fig-\nure 2 shows the photoinduced Faradayrotation (PFR) as\na function of applied magnetic field, for CW excitation\nwith intensity p= 400 mW/cm−2, atT= 12 K. The\nPFR signal has all the characteristics expected for pho-\ntoinduced magnetic polarons: no signal at zero magnetic\nfield, because polarons are radomly oriented and produce\nzero magnetization, and a very rapid increase when a\nmagnetic field is applied, due to the complete alignment\nofthepolaronensemblealongthefieldduetotheZeeman\n∗Permanent address: Ioffe Institute, 194021 St. Petersburg, Russia\nFIG. 1: Scheme of the setup for measuring the time-resolved\nphotoinduced Faraday rotation angle. The linearly polariz ed\nprobe pulse arrives at the sample when a time ∆ thas elapsed\nafter the arrival of the pump pulse. We measured ∆ θF, the\nFaraday rotation of the probe, induced by the pump illumi-\nnation.\ntorque acting on particles with a largemagnetic moment.\nFig. 2 shows that when the applied magnetic field is in-\ncreased beyond 1T, the PFR effect gradually decreases\ntowards zero; this is because large magnetic fields pro-\ngressively force the EuSe lattice spins into ferromagnetic\norder, and ultimately the formation of spin polarons is\nquenched. Fig. 3 shows that the PFR signal vanishes if\nthe excitation photon energy is less than the bandgap,\nand increases abruptly at the bandgap energy, demon-\nstrating that the PFR effect is associated with photogen-\nerated conduction band electrons.\nBecause the photoexcited electrons are bound to very\nheavyphotoexcitedholes[4], thespin polaronsareimmo-\nbile and are confined to a layer below the surface of the\ncrystal whose thickness equals the penetration depth of\nthe excitation light, 1 /α, whereα∼=15µm−1is the EuSe\nabsorption coefficient for the excitation photon energy\n[4]. From elementary kinetics, the steady-state density\nof photoinduced polarons is given by[5]\nnPol=χpατPol\nhν≤pατPol\nhν, (1)\nwhereχ≤1 is the quantum efficiency for spin polaron\nphotogeneration, and τPolis the spin polaron lifetime.DRAFT2\nFIG. 2: Photoinduced Faraday rotation in EuSe. The max-\nimum PFR is indicated by ∆ θSAT\nF.\nFIG. 3: PFR excitation spectrum.\nThe spin polaron lifetime was deduced from the depen-\ndenceofthePFRamplitudeonthemodulationfrequency\nof the excitation light, shown in Fig. 4, following the pro-\ncedure described in 6, giving τPol=1.6µsec. The mean\ndistance between polarons, d, in units of the EuSe lattice\nconstant, is therefore\nd\na≥2\na/parenleftbigg3\n4πhν\npατPol/parenrightbigg1/3\n(2)\nSubstituting all parameters into (2), we obtain dPol≥\n36a. Thus,dis at least one order of magnitude greater\nthan the typical radius of a spin polaron in europium\nchalcogenide [8], RPol∼3. Being so far apart, the po-\nlarons are non-interacting. Moreover, above the N´ eel\ntemperature the spin polarons orientation floats freely,\nFIG. 4: PFR as a function of the pump modulation fre-\nquency.\nFIG. 5: PFR as a function of internal magnetic field. The\nratioχ⊥/χ||, used to compute the internal magnetic field, is\nshown in the inset, as a function of temperature.\ntherefore they form a superparamagnetic gas, whose\nmagnetization dependence on magnetic field and tem-\nperature obeys a Langevin function [9]. Because PFR is\nproportionalto the magnetization [10], the photoinduced\nFaraday rotation angle will be given by\n∆θF= ∆θSAT\nFL/parenleftbiggµPolB\nkBT/parenrightbigg\n(3)\nThe magnitude of the magnetic moment of the spin\npolaron at a given temperature can be determined ac-\ncurately by fitting the experimental curve with (3), be-\ncause it is the sole adjustable parameter determining the\nsharpness of the step. However, we must first convert the\napplied magnetic field into internal one, which is smallerDRAFT3\nFIG. 6: Magnetic moment of a photoinduced polaron as a\nfunction of temperature. Full circles, empty circles and in -\nverted triangles correspond to probe wavelength of 665, 730\nand 760 nm, respectively. The dashed line shows the para-\nmagnetic approximation, with JXf= 0.075 eV obtained from\nthe linear fit shown in the inset.\nthan the former due to the demagnetizing field [11]. For\ntheFaradaygeometryusedhere,wherethemagneticfield\nis normal to a very thin epitaxial layer, the ratio of the\ninternal magnetic field to the applied one is equal to the\nratio of the magnetic susceptibility measured when the\napplied field is perpendicular to the layer, ( χ⊥), to the\nsusceptibility measured when the applied field is parallel\nto it (χ||). The susceptibilities χ⊥andχ||were mea-\nsured using a SQUID magnetometer, which had a mag-\nnetic moment resolution better than 10−11Am2. Figure\n5 shows the measured PFR as a function of internal mag-\nnetic field. The inset shows the measured ratio between\nsusceptibilities χ⊥andχ/bardbl, used to convert the applied\nmagnetic field into internal one, above the N´ eel temper-\nature. The solid line in Fig. 5 shows the fit of the theory,\nusing (3), which yields the magnetic moment of the spin\npolaron,µPol= 5,590 Bohr magnetons ( µB). This mag-\nnetic moment is 10 times greater than the giant photoin-\nduced spin polarons known so far [5, 6], which allows us\nto categorize the spin polarons discovered as supergiant.\nThe magnetic moment of the spin polaron as a function\nof temperature is shown in Fig. 6. Increasing the tem-\nperature quenches spin polaron formation, due to the de-\ncreasing magnetic susceptibility of the lattice spins. At\ntemperatures much higher than the N´ eel temperature,\nandµEuB/kBT≪1, the magnetization of lattice spins\ncanbewelldescribedbytheparamagneticapproximation\nM=NµEuS+1\n3SµEuB\nkBT. (4)\nwhereN= 4/a3is the density of Eu sites in the crystal\nandµEu=gµBSis the magnetic moment of an Eu atom,whereg= 2,S= 7/2. The exchange interactionbetween\nthe photoexcited electron, whose wavefunction is ψ(r),\nandthelatticespins,isdescribedbyaneffectivemagnetic\nfield,BXf, acting on the lattice spins [12]\nBXf=JXfS\nNµEu|ψ(r)|2, (5)\nwhereJXfistheexchangeinteractionintegralbetweenthe\nelectronformingthespinpolaronandthelatticespins[8].\nSubstituting (5) in (4), and integrating in the whole vol-\nume, the high temperature magnetic moment of a spin\npolaronin the paramagneticcrystal approximationis ob-\ntained\nµPol(T) =gµBS(S+1)JXf\n3kBT, (6)\nAfitofthehightemperaturetail, T >40K,ofµPolversus\nTwith (6), whereby JXfis the only adjustable parameter,\nis shown in the inset of Fig. 6, and gives JXf= 75 meV.\nThisJXfvalue is about the same as measured for EuTe,\nmeaning that the mean exchange field generated by the\nphotoexcited electron in EuSe is about the same as for\nEuTe, i.e. about 1 Tesla [8]. Fig. 6 shows that when the\ntemperature is decreased below 20K and approaches the\nN´ eel temperature, µPolincreases much faster than the\nparamagnetic approximation, and can reach almost an\norder of magnitude greater thanµPolthat would be ob-\nserved if the exchange interaction was switched off (the\nparamagnetic approximation). This is in sharp contrast\nto EuTe, where µPolis always smaller than the param-\nagnetic limit [6]. The reason for the very large increase\nin EuSe is the near equal absolute values of the first and\nsecond neighbor exchange interaction constants, J1and\nJ2. Whereas J1>0 favors ferromagnetism, J2<0 fa-\nvors antiferromagnetism, which predominates only be-\ncause|J2|is marginally larger than |J1|. This also im-\nplies than around the N´ eel temperature only a very small\ninternal magnetic field, of about 0.1 T is sufficient to\npromote a phase transition into the ferromagnetic state,\nas demonstrated experimentally in Ref. 13. Bearing in\nmind that within a spin polaron the effective magnetic\nfield acting on the spins is about 1 T, this implies that\nthe within the polaron the crystal lattice attains ferro-\nmagnetic alignment, which explains why the magnetic\nmoment of the polaron is so large in EuSe. Such reason-\ning is supported quantitatively by the Weiss mean field\napproximation (MFA) with first and second neighbor in-\nteraction at T= 0K, according to which in the AFM-II\nphase the magnetic field to induce ferromagnetic align-\nment is given by [8, 12, 14]\nBSAT=24|J1+J2|S\ngµB, (7)\nand in the range B < B SATthe magnetization is given\nbyM=NµEuB/B SAT. Using (5) and (7), and integrat-\ningMin the volume, the magnetic moment of the spinDRAFT4\nFIG. 7: PFR as a function of pump excitation power.\npolaron in the mean field approximation is obtained\nµMFA\nPol(T= 0K) =JXf\n24|J1+J2|gµB. (8)\nTable I displays the result produced by formula (8) for\nEuTe and EuSe. The calculated values agree by order of\nmagnitude with the measured values, and demonstrate\nthat the very large µPolin EuSe is due to the near can-\ncellation of the ferro and antiferromagnetic lattice inter-\nactions.\nTABLE I: EuSe and EuTe parameters, µPolmeasured at T=\n5 K and estimated using the MFA for T= 0 K.\nµPol(µB)\nJ1(K)J2(K)JXf(eV)Experiment MFA\nEuTe 0.043 -0.15 0.083 600 750\nRef. 15 12 5\nEuSe 0.29 -0.30 0.075 5,590 7,300\nRef. 16 this work\nThe MFA does not provide any information about the\ninternal structure of the spin polaron. A more accurate\ndescription of the spin polaron, and a description of its\ninternal structure, is provided by a self-consistent calcu-\nlation for an AFM-II type system at T= 0 K [12]. In\nperforming such a calculation for EuSe, the input param-\neters were a conduction band effective mass m∗= 0.3m0\n[17], a dielectric constant ε= 9.4 [7], as well as the pa-\nrameters given in Table I. The self-consistent calculation\nproduces a spin polaron containing a large ferromagnetic\ncore of radius 2.8 lattice parameters, containing 370 fer-\nromagnetically aligned Eu atoms, with a calculated spin\npolaron of magnetic moment of µPol= 4,920µB.\nFigure 7 shows ∆ θSAT\nF, and the corresponding polaron\npopulation, deduced using the Verdet constant as de-\nFIG. 8: PFR as a function of the delay, ∆ t, between pump\nand probe pulses.\nscribed in Ref.10, as a function of excitation pump power\ndensity. The solid line is a fit using (1) with χas the\nonly fitting parameter, which yields χ= 0.053. In dra-\nmatic contrast to EuTe, which shows a saturation of the\npolaron population, in EuSe the population grows lin-\nearly with excitation even at concentrations far above\nexpected residual deffect concentrations, indicating that\nthe photoinduced polarons are intrinsic, and therefore\nis should be possible to fully magnetize the layer pene-\ntrated by light simply by using enough excitation power.\nThe maximum pump intensity shown in Fig.7 generated\nabout 2×1017cm−3spin polarons, and taking into ac-\ncount their ferromagnetic core radius of about 3 a, this\ncorresponds to a magnetization of 5 ×10−3of the satura-\ntion value. We could not exploit greater pump intensities\nwithout heating of the sample, because we used an opti-\ncal chopper with a high (50%) duty cycle.\nFigure 8 shows the PFR signal as a function of the\ndelay, ∆t, between the pump and probe pulses (Fig.1).\nBefore the arrival of the pump pulse (negative delay), no\nPFR signal is detected, however, after the arrival of the\npump pulse the PFR signal grows exponentially with a\ncharacteristic rise time of 62 ps, which classifies the pho-\ntomagnetization process as ultrafast. After achieving the\nmaximum, the PFR remained approximately constant\nwithin the maximum delay available of our experiment,\nof a few nanoseconds, which is explained by the long life-\ntime of the polarons, τPol= 1.6µs, reported above.\nIn conclusion, we have demonstrated that light can be\nusedtoconvertazeromagnetizationstateofaEuSecrys-\ntal into a completely polarized ferromagnetic state in the\nultrafast time scale, through the photogeneration of su-\npergiantintrinsicspinpolarons. Thismagetizationmech-\nanism is made possible because the exchange interaction\nbetween lattice spins in EuSe contains both ferromag-DRAFT5\nnetic and antiferromagnetic components, which nearly\ncancel each other out.\nThis work was supported by the Brazilian agencies\nCNPq (Project 304685/2010-0) and FAPESP (Project\n2016/24125-5).\n[1] J. Xie, H. Qin, Y. Hao, B.Cheng, W. Liu, L. Liu, S. Ren,\nG. Zhou, Z. Ji, and J. Hu, Sci. Rep. 7, 45642 (2017).\n[2] T. Satoh, R. Iida, T. Higuchi, Y. Fujii, A. Koreeda,\nH.Ueda, T. Shimura, K. Kuroda, V. I. Butrim, and B. A.\nIvanov, Nature Communications 8, 638 (2017).\n[3] I. Radu et al, Nature 472, 205 (2011).\n[4] A. B. Henriques, A. Wierts, M. A. Manfrini,\nG. Springholz, P. H. O. Rappl, E. Abramof, and A. Y.\nUeta, Phys. Rev. B 72, 155337 (2005).\n[5] A. B. Henriques, A. R. Naupa, P. A. Usachev, V. V.\nPavlov, P. H. O. Rappl, and E. Abramof, Phys. Rev. B\n95, 045205 (2017).\n[6] A. B. Henriques, G. D. Galgano, P. H. O. Rappl, and\nE. Abramof, Phys. Rev. B 93, 201201 (2016).[7] A. Mauger and C. Godart, Phys. Rep. 141, 51 (1986).\n[8] A. B. Henriques, G. D. Galgano, E. Abramof, B. Diaz,\nand P. H. O. Rappl, Appl. Phys. Lett. 99, 091906 (2011).\n[9] C. P. Bean and J. D. Livingston, J. Appl. Phys. 30, 120S\n(1959).\n[10] A. B. Henriques and P. A. Usachev, Phys. Rev. B 96,\n195210 (2017).\n[11] S. Blundell, Magnetism in Condensed Matter (Oxford\nUniversity Press, 2001), Appendix D.\n[12] A. B. Henriques, F. C. D. Moraes, G. D. Galgano, A. J.\nMeaney, P. C. M. Christianen, J. C. Maan, E. Abramof,\nand P. H. O. Rappl, Phys. Rev. B 90, 165202 (2014).\n[13] R. T. Lechner, G. Springholz, T. U. Sch¨ ulli, J. Stangl,\nT. Schwarzl, and G. Bauer, Phys. Rev. Lett. 94, 157201\n(2005).\n[14] W. S¨ ollinger, W. Heiss, R. T. Lechner, K. Rumpf,\nP. Granitzer, H. Krenn, and G. Springholz, Phys. Rev.\nB81, 155213 (2010).\n[15] P. Wachter, C R C Critical Reviews in Solid State Sci-\nences3, 189 (1972).\n[16] X. Gratens et al, unpublished (2018).\n[17] S. J. Cho, Phys. Rev. B 1, 4589 (1970)." }, { "title": "1011.2788v1.Spin_Torque_Ferromagnetic_Resonance_Induced_by_the_Spin_Hall_Effect.pdf", "content": " 1Spin Torque Ferromagnetic Resonance Induced by the Spin Hall Effect \nLuqiao Liu, Takahiro Moriyama, D. C. Ralph, and R. A. Buhrman \nCornell University, Ithaca, New York, 14853 \n \nABSTRACT \nWe demonstrate that the spin Hall effect in a thin film with str ong spin-orbit scattering \ncan excite magnetic precession in an adjacent ferromagnetic film. Th e flow of alternating current \nthrough a Pt/NiFe bilayer generates an oscillatin g transverse spin current in the Pt, and the \nresultant transfer of spin angular momentum to the NiFe induces ferromagnetic resonance (FMR) \ndynamics. The Oersted field from the current also generates an FMR signa l but with a different \nsymmetry. The ratio of these two signals allows a quantitative determinati on of the spin current \nand the spin Hall angle. \n 2The spin Hall effect (SHE), the conversion of a longitudinal charge current density JC \ninto a transverse spin current density / 2SJe= , originates from spin-orbit scattering [1-4], \nwhereby conduction electrons with opposite spin orientations in a nonmagnetic metal [5] or \nsemiconductor [6] are deflected in opposite di rections. The SHE has attracted widespread \ninterest because it can generate pure spin currents from a nonmagnetic source, a phenomenon \nthat could find important applications in future spintronic devices. Seve ral techniques [5, 7, 8] \nhave been developed to determine the magnitude of the SHE, which is generally characterized by \nthe spin Hall angle, θSH = JS/JC. For thin-film Pt, estimates of θSH obtained using different \napproaches differ by more than an order of magnit ude [8-10], but already there have been efforts \nto utilize the spin current that arises from the SHE, first to tune the damping coefficient in a \nferromagnetic metal [8], and, most recently, to indu ce a spin wave oscillation in a ferrimagnetic \ninsulator having small damping [11]. Here we show that the SHE can also be used to excite \ndynamics in an ordinary metallic ferromagnet. Our experiment also allows a quantitative \ndetermination of the SHE strength that is self-calibrated, as explained below, enabling \nmeasurements of the spin currents generated by th e SHE with small experimental uncertainties. \nWe study Pt/Permalloy bilayer films with a microwave-frequency (RF) charge current \napplied in the film plane (Permalloy = Py = Ni 81Fe19). An oscillating transv erse spin current is \ngenerated in the Pt by the SHE and injected into the adjacent Py (Fig. 1(a)), thereby exerting an \noscillating spin torque (ST) on the Py th at induces magnetization precession. When the \nfrequency and field bias satisfy the FMR condition for the Py, st rong resonant precession results \nin a significant oscillation of the bilayer resistance due to the anisotropic magnetoresistance \n(AMR) of the Py. This generates a DC voltage signal across the sample from the mixing of the \nRF current and the oscillating re sistance, similar to the signal th at arises from ST induced FMR 3in spin valves and magnetic tunne l junctions [12-15]. The resonan ce properties enable a direct \nquantitative measure of the spin current absorbed by the Py. \nOur measurement setup is shown in Fig. 1(c). Pt/Py bilayers were grown by DC \nmagnetron sputter deposition. The in dividual layer thicknesses we re 4-15 nm, with specific \nvalues stated below. The star ting material for the Pt was 99.95% pure. Highly resistive Ta (1 nm) \nwas employed as the capping layer to preven t oxidation of the Py. The bilayers were \nsubsequently patterned into microstrips usi ng photolithography and ion milling. The samples’ \nwidths ranged from 1 to 20 μm and the lengths from 3 to 250 μm. By using a bias tee, we were \nable to apply a microwave current to our sample and at the same time me asure the DC voltage. A \nsweeping magnetic field Hext was applied in the film plane, with the angle θ between Hext and \nmicrostrip kept at 45° unless otherwise indi cated. The output power of the microwave signal \ngenerator was varied from 0 to 20 dBm and th e measured DC voltage was proportional to the \napplied power, indicating that the induced precession was in the small angle regime. All the \nmeasurements we present were performed at room temperature with a power of 10 dBm. \nWe model the motion of the Py magnetic moment ˆm by the Landau-Lifshitz-Gilbert \nequation containing the ST term [16]: \n,\n0ˆˆˆˆ ˆ ˆ ˆ ˆ ()2eff S RF RF\nSdm dmmH m J m m mHdt dt e M tγα γ σ γμ=− × + × + × × − ×GG =. (1) \nHere γ is the gyromagnetic ratio, α is the Gilbert damping coefficient, μ0 is the permeability in \nvacuum, Ms is the saturation magnetization of Py, t is the thickness of the Py layer, ,/2SR FJe= \nrepresents the oscillating spin curr ent density injected into Py, HRF is the Oersted field \ngenerated by the RF current, Heff is the sum of Hext and the out-of-plane demagnetization field \n 4πMeff, and ˆσ is the direction of the injected spin moment. The third and fourth terms on the 4right hand side of Eq. (1) are the result of in-plane spin torque and the out-of-plane torque due \nto the Oersted field, respectiv ely (Fig. 1(a)). The magnetic-re sonance mixing signal in response \nto a combination of in-plane a nd out-of-plane torques has been calculated in the context of ST-\ndriven FMR [14, 15], which we can translate to our notation as: \nVmix=−1\n4dR\ndθγIRFcosθ\nΔ2πdf/dH()Hext=H0SFS(Hext)+AFA(Hext) [] , (2) \nwhere FSHext() =Δ2/Δ2+Hext−H0 ()2⎡⎣⎤⎦ is a symmetric Lorentzian function centered at the \nfield H0 with linewidth Δ, FAHext() =FSHext() Hext−H0 () /Δ is an antisymmetric Lorentzian, \n(),0/2SR F s SJ e M t μ == , A=HRF1+4πMeff/Hext ()⎡⎣⎤⎦1/2, R is the resistance of the stripline, IRF \nis the microwave current through the stripline, and f is the resonance frequency. We therefore \nexpect the resonance signal to consist of two pa rts, a symmetric Lorentzian peak proportional to \nthe spin current density and an an tisymmetric peak proportional to HRF. \nThe Oersted field HRF can be calculated from the geometry of the sample. Since the \nmicrowave skin depth is much greater than the Py thickness the current de nsity in the Py should \nbe spatially uniform, and in this case the Oersted field from the charge current density in the Py \nshould produce no net torque on the Py [see Fig. 1(b)]. The Oersted field can therefore be \ncalculated entirely from the current density JC,RF in the Pt layer. The microstrip width is much \nlarger than the Pt thickness, so the sample can be approximated as an infinitely wide conducting \nplate and the Oersted field determined by Ampère's law, HRF=JC,RFd/2, where d is the Pt \nthickness. We checked HRF by numerical integration and the di fference is less than 0.1% from \nthe infinite plate approximation. Using this result , the ratio of the spin current density entering \nthe Py to the charge current density in the Pt can then be determined quantitatively in a simple 5way from the ratio of the symmetric and anti symmetric components of the resonance curve \n()1/2, 0\neff ext\n,14 /SR F S\nCR FJ eM t dSMHJAμπ =+ ⎡ ⎤⎣ ⎦=. ( 3 ) \nAll of the parameters entering Eq . (3) are either fundamental cons tants or quantities that can be \nmeasured directly, so this expression allows a measurement of JS,RF/JC,RF with small \nexperimental uncertainties. The measurement is se lf-calibrated in the sense that the strength of \nthe torque from the spin current is m easured relative to the torque from HRF, which can be \ncalculated easily from the geometry of the sample. \nAn additional contribution to the DC voltage can arise from spin pumping by the \nprecessing moment in combination w ith the inverse SHE in the Pt la yer, as observed in Ref. 10. \nHowever, this effect is second order in θSH in our geometry and we calculate that it should \ncontribute a negligible voltage, about two orders of magnitude smaller than the signals we \nmeasure. \nFigure 2(a) shows the ST-FMR signals m easured on a Pt(6)/Py(4) (thicknesses in \nnanometers) sample for f = 5-10 GHz. As expected from Eq. (2), the resonance peak shapes can \nbe very well fit by the sum of symmetric and antisymmetric Lorentzian curves with the same linewidth for a given f (fits are shown as lines in Fig. 2(a)). The fact that the symmetric peak \nchanges its sign when H\next is reversed (inset of Fig. 2(a)) ag rees with the form of spin torque \nˆˆˆST mm τσ∝× ×G given in Eq. (1), and excludes the possi bility that the signal is due to an \nunbalanced perpendicular Oerste d field torque, in direction ˆˆRF mH⊥× , which would yield \nsymmetric peaks with the same sign for opposite Hext. The resonant peak positions are \nsummarized in Fig. 2(b), and agree well with the Kittel formula \n() ( )1/2\n00 e f f /2 4 fH H Mγπ π=+ ⎡⎤⎣⎦. From a one-parameter fit to the resonance frequencies we 6determine that the demagnetization field 4πMeff= 0.805 ± 0.005 T for the Pt(6)/Py(4) bilayers. \nWe have also measured th e saturation magnetization MS = 6.4 × 105 A/m in test samples [17]. \nTo verify the SHE origin of field-symmetr ic components of the FMR signals, we have \nstudied several different types of control samples. In Fig. 2(c) we compare the FMR signals \nmeasured at 8 GHz for a Pt(15)/Py(15) and a Pt (6)/Py(4) sample. The signal for the Pt(6)/Py(4) \nsample contains a sizable field-symmetric component, with S/A = 0.63. Due to the increased \nthicknesses of the two layers, we expect from Eq. (3) that S/A for the Pt (15)/Py (15) should be \ngreatly reduced, approximately ∝1/td if the spin Hall currents in the two samples are similar. \nS/A for the Pt(15)/Py(15) is very small, S/A = 0.08 ± 0.05, near the noise floor for the fits of the \nsymmetric component (the uncertainty reflects the standard deviation over five samples \nmeasured). The difference between the change in the S/A ratio expected from Eq. (3) (a factor of \n11.2, taking into account a small change in 4πMeff) and the measured reduction by a factor 8.0 \nmay be associated with a change in the magnit ude of the spin Hall current generated by the \ndifferent thicknesses of the Pt films when this thickness is comparable to the spin diffusion length (see below). \nWe also studied control samples with the laye rs Cu(6)/Py(4) and 4 nm of Py alone, with \nresults as shown in Fig. 2(d). The Cu/Py bila yer sample gives a purely antisymmetric signal, \nindicating that only the Oersted-fi eld contribution is present, as expected because of the very \nsmall SHE in Cu in comparison to that in Pt . For the Py(4) sample, we would expect no \nresonance signal at all, since ther e is no SHE and as noted above if the current density in the Py \nis uniform there should also be no net effect of the Oersted field on the Py dynamics. However, \nwe do observe a very small, purely antisymmetric signal in the 4 nm Py sa mple. We suspect that \nthis may arise from an Oersted field due to non-unif orm current flow at the ends of the Py due to 7the electrode contacts. The lack of field-symmet ric components in the re sonance curves for the \ncontrol samples provides strong su pport that the symmetric component we observe in Pt(6)/Py(4) \ndoes indeed arise from the SHE in the Pt. \nWith 4πMeff and MS determined , we can use Eq. (3) and the measured values of S/A to \ncalculate JS,RF/JC,RF. The results are shown in Fig. 2(e) for the resonance curves spanning 5-\n10 GHz shown in Fig. 2(a). We find JS,RF/JC,RF = 0.056 ± 0.005 for Pt(6)/Py(4). We \nmeasured more than ten Pt(6)/Py(4) samples w ith different lateral dimensions and the total \nvariation of JS,RF/JC,RF was < 15%. The dominant experimental uncertainty [and the small \nvariation with Hext visible in Fig. 2(e)] may be associated with Oersted fields from non-uniform \ncurrents at the sample ends, as noted above for the single-layer Py sample. Note that according to Eq. (2) S/A should not depend upon the angle of the a pplied DC field, as confirmed by the \nresults shown in Fig. 2(f). \nAs an independent check we also employed an alternative method for determining the \nspin current density absorbed by the Py layer, by measuring the FMR linewidth \nΔ as a function \nof DC current, similar to the technique introduced in Ref. 8. According to the theory of ST, a \nDC spin current IS,DC will increase or decrease the effective magnetic damping and hence Δ, \ndepending upon its relative or ientation with respect to the magnetic moment: [18] \neff 02s i n\n(2 ) 2S\next SJ f\nHM M t eπθαγπ μ⎛⎞Δ= +⎜⎟+ ⎝⎠= (4) \nOur results obtained with a Pt(6)/Py(4) sample ~1 μm wide are shown in Fig. 3. The measured \ndamping coefficient at zero current ( α ≈ 0.028) is significantly higher than that measured in a \nspin valve nanopillar sample ha ving a 4 nm Py free layer ( α≈ 0.01 ) [19]. This can be \nexplained by the spin pumping e ffect previously observed in the Py/Pt system [20, 21]. For a 8negative applied field ( Hext applied -135° from the current di rection in the microstrip), the \nlinewidth is broadened when I DC ramps from -0.7 mA to 0.7 mA; while for a positive field \n(Hext applied 45° from the current direction), th e trend is the opposite. By fitting the data \nshown in Fig. 3, and calculating the charge cu rrent density in the Pt using the measured \nresistivities 20Pt cm ρμ=Ω and 45Py cm ρμ=Ω , we find 10/ (0.9 0.012) 10cJα−Δ= ± × (A/cm2)-1. \nWith Eq. (4), this yields / 0.048 0.007scJJ =± for Pt(6)/Py(4), which agrees well with the \nvalue 0.056 ± 0.005 determined from the FMR lineshape. \nOur experiments yield values for JS/JC, the ratio of the spin current density (in units of \ncharge) absorbed by the Py to the charge current density in the Pt film. For many applications, \nthis is the figure of merit of direct interest. Ho wever, for comparing to other experiments, it is \nalso of interest to determine the spin Hall angle θSH, the ratio of the spin current density inside \nbulk Pt to the charge current density. For a perf ectly transparent Pt/Py interface and for a Pt \nlayer much thicker than the spin diffusion length λsf, the quantities JS/JC and θSH should be \nequal. However, because our Pt/Py interface is li kely not perfectly transparent, and because our \nPt layers likely do not have thicknesses \u0015λsf, our results may underestimate the transverse \nspin current density appropriate to bulk Pt. Therefore, our measurements imply a lower bound, \nθSH > 0.056 ± 0.005 for our Pt material. In the lim it of a transparent Pt/P y interface, for which \nthere should be no spin accumulation transverse to the Py moment at the interface, we calculate \nusing drift-diffusion theory [22] that the spin Hall current density in a Pt film of thickness d \nshould be reduced from the bulk value by JSd()/JS∞()=1−sech d/λsf(). Using this \nexpression, our best estimate, based on compar ison between the Pt(15)/Py(15) and Pt(6)/Py(4) \nsamples is that 3sf nm λ≈ , and we can set an upper bound of 6sf nm λ< , lower than the low 9temperature value measured previously [23]. This gives a best estimate of 0.076SHθ= , and \nbounds 0.056 0.005 0.16SHθ ±< < , again in the limit of a tr ansparent Pt/Py interface. \nWe mentioned above that pr evious measurements of θSH in Pt have differed by over an \norder of magnitude. Kimura et al. [9], using a Pt/Cu/Py lateral nonlocal geometry reported \nθSH= 0.0037. However, their 4-nm-thick Pt wires are in contact to 80-nm -thick Cu wires. We \nbelieve that the Cu likely shunted the charge current flowing in the Pt, resulting in a large \nunderestimation of θSH. Ando et al. [8] by measuring magnetic damp ing in Pt/Py versus current, \nreported JS/JC = 0.03 and estimated 0.08SHθ= . We have shown that a technique closely \nrelated to the method of Ref. 8 gives result s that agree with our FMR method, although we \ndiffer with Ref. 8 regarding the form of our Eq. (4) and the drift-di ffusion analysis. Mosendz et \nal. [10], using a technique based on spin pump ing together with the inverse SHE, reported θSH= \n0.0067, later refined to θSH = 0.013 [24]. This result relied on an assumption that λsf= 10 nm \nfor Pt. Their value for θSHwould be 3 times larger, and in much better accord with our value, \nusing our estimate that λsf= 3 nm. \nIn summary, we demonstrate th at spin current generated by the SHE in a Pt film can be \nused to excite spin-torque FMR in an adjacen t metallic ferromagnet (Py) thin film. This \ntechnique allows a straightforward determination of the efficiency of spin current generation, \nJs/Jc (the spin current density absorbed by the Py di vided by the charge curre nt density in the Pt), \nthat is self-calibrated, in that the torque due to the spin current can be measured relative to the \ntorque from the Oersted field generated by the sa me charge current density in the Pt layer. We \nfind Js/Jc = 0.056 ± 0.005 for Pt(6)/Py(4), implying θSH > 0.056 for bulk Pt . This simple \ntechnique is an excellent soluti on for the quantitative measuremen t of the SHE efficiency in any 10metallic film that can be produced as part of a ferromagnetic/non-magnetic metal bilayer. The \nrelatively large efficiency of spin current generation that we observe for Py/Pt is promising for applications which might utilize the SH E to manipulate ferromagnet dynamics. \nThis research was supported in part by th e Army Research Office, and by the NSF-NSEC \nprogram through the Cornell Center for Nanoscale Sy stems. This work was performed in part at \nthe Cornell NanoScale Facility, which is supported by the NSF through the National \nNanofabrication Infrastructure Network and bene fitted from the use of the facilities of the \nCornell Center for Materials Research , supported by the NSF-MRSEC program. 11REFERENCES \n1. J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999). \n2. S. F. Zhang, Phys. Rev. Lett. 85, 393 (2000). \n3. J. Sinova et al. , Phys. Rev. Lett. 92, 126603 (2004). \n4. S. Murakami, N. Nagaosa and S. C. Zhang, Science 301, 1348 (2003). \n5. S. O. Valenzuela and M. Tinkham, Nature 442, 176 (2006). \n6. Y. K. Kato et al. , Science 306, 1910 (2004). \n7. E. Saitoh et al. , Appl. Phys. Lett. 88, 182509 (2006). \n8. K. Ando et al. , Phys. Rev. Lett. 101, 036601 (2008). \n9. T. Kimura et al. , Phys. Rev. Lett. 98, 156601 (2007). \n10. O. Mosendz et al. , Phys. Rev. Lett. 104, 046601 (2010). \n11. Y. Kajiwara et al. , Nature 464, 262 (2010). \n12. A. A. Tulapurkar et al. , Nature 438, 339 (2005). \n13. J. C. Sankey et al. , Phys. Rev. Lett. 96, 227601 (2006). \n14. J. C. Sankey et al. , Nat. Phys. 4, 67 (2008). \n15. H. Kubota et al. , Nat. Phys. 4, 37 (2008). \n16. J. C. Slonczewski, J. Magn. Magn. Mater. 159, L1 (1996). \n17. I. N. Krivorotov et al. , Science 307, 228 (2005). \n18. S. Petit et al. , Phys. Rev. Lett. 98, 077203 (2007). \n19. G. D. Fuchs et al. , Appl. Phys. Lett. 91, 062507 (2007). \n20. S. Mizukami, Y. Ando and T. Miyazaki, J. Magn. Magn. Mater. 239, 42 (2002). \n21. Y. Tserkovnyak, A. Brataas and G. E. W. Bauer, Phys. Rev. Lett. 88, 117601 (2002). \n22. P. C. Vanson, H. Vankempen and P. Wyder, Phys. Rev. Lett. 58, 2271 (1987). \n23. H. Kurt et al. , Appl. Phys. Lett. 81, 4787 (2002). \n24. A. Hoffman, private communication. \n \n 12FIGURE CAPTIONS \nFig. 1 (color online): (a) Schematic of a Pt/Py bi layer thin film illustrating the spin transfer \ntorque τSTT, the torque τH induced by the Oersted field HRF , and the direction of the damping \ntorque τα. θ denotes the angle betw een the magnetization M and the microstrip. Hext is the applied \nexternal field. (b) Left side view of the Pt/P y system, with the solid line showing the Oersted \nfield generated by the current flowing just in the Py layer, which should produce no net effect on \nthe Py AMR. (c) Schematic circuit for the ST-FMR measurement. Fig. 2 (color online): (a) Spectra of ST-F MR on a Pt(6)/Py(4) sample measured under \nfrequencies of 5-10 GHz. The sample dimension is 20 μm wide × 110 μm long. Inset: ST-FMR \nspectrum of 8 GHz for both positive and negative H\next. (b) Resonance frequency f as a function \nof the resonant field H0. The solid curve represents a fit to the Kittel formula. (c) FMR spectra \nmeasured for two Pt/Py bilayer samples, with fits to Eq. (2). The data were taken at 8 GHz. (d) \nFMR spectra ( f = 8 GHz) on the Pt(6)/Py(4) sample (blue triangles) as well as control samples \nconsisting of Cu(6)/Py(4) (red circles) and Py(4) (black squares). (e) JS,RF/JC,RF values \ndetermined from the FMR analysis [Eq. (3)] at different f. (f) FMR signals measured for different \nexternal field angles θ (f = 8 GHz). The mixing voltages Vmix are normalized and offset to enable \ncomparison of the lineshapes. Fig.3 (color online): The change of the FMR linewidth (left y axis) and Gilbert damping coefficient (right y axis) as a function of I\nDC for two orientations of the Py magnetization relative \nto the current direction. The data are taken at f = 8 GHz. 13\n \n 14 \n0 1000 2000-2002040\n Vmix (μV)\nHext (Oe)5 GHz\n6 GHz\n7 GHz\n8 GHz\n9 GHz\n10 GHz\n \n0 500 1000 15000246810\n f (GHz)\nH0 (Oe)\n-600 -300 0 300 6000.00.81.62.43.2 \n \n75o60o45oNormalized V mix\nHext - H0 (Oe)θ=15o\n30o\n400 800 12000.000.020.040.06\n Js / Jc\nH0 (Oe)0 1000 2000-20-1001020\n Vmix (μV)\nHext (Oe) Pt (6) / Py (4)\n Cu (6) /Py (4)\n Py (4)\n0 1000 2000-40-2002040\n Vmix (μV)\nHext (Oe)Pt (15) / Py (15)\nPt (6)/Py (4)(b)\n(f)(c) (d)\n(e)(a) 15 \n-0.8 -0.4 0.0 0.4 0.87578818487\n2.72.82.93.0 \n positive field\n negative fieldlinewidth (Oe)\nIDC (mA)X 10-2\n α \n " }, { "title": "2311.05868v1.Observation_of_the_out_of_plane_orbital_antidamping_like_torque.pdf", "content": "Observation of the out-of-plane orbital antidamping -like torque \nZeyang Gong1†, Fu Liu1†, Xinhong Guo1, Changjun Jiang1* \n1 Key Laboratory for Magnetism and Magnetic Materials, Ministry of Education, Lanzhou \nUniversity, Lanzhou 730000, China \n \n†These authors contributed equally to this work. \nCorresponding author. E -mail address: *jiangchj@lzu.edu.cn \n \n \nAbstract \nThe out -of-plane antidamping -like orbital torque fosters great hope for high -\nefficiency spintronic devices. Here w e report experimentally the observation of out -of-\nplane antidamping -like torque that could be generated by z-polarized orbital current in \nferromagnetic -metal/oxidized Cu bilayers , which is presented unambiguously by t he \nmagnetic field angle dependence of spin-torque ferromagnetic resonance signal . The \noxidized Cu thickness dependence of orbital torque ratios highlights the interfacial \neffect would be responsible for the generation of orbital current . Besides that, the \noxidized Cu thickness dependence of damping parameter further proves the observation \nof antidamping -like torque. This result contributes to enriching the orbital -related \ntheory of the generation mechanism of the orbital to rque. \n \n 1 Introduction \nOrbital current, a flow of orbital angular momentum (OAM) , have attract ed wide \ninterest in the recent past .1-4 Orbital current is defined as the directional flow of carriers \ncarrying OAM , which carriers angular momenta polarized in different directions \naccording to the different direction of their own OAM .5 Generally, there are two \nmechanisms giving rise to the generation of orbital current, orbital Hall effect (OHE)6-\n7 and orbital Rashba effect (ORE) .8-9 OHE refers to a longitudinal electric field induc ing \nan excitation of orbital current in a transverse direction. While the ORE allows the \ngeneration of chiral OAM texture in k space via OAM dependence of energy splitting \nanalogous to spin Rashba effect (SRE). Notablely, unlike SRE, the ORE can still exist \nwithout spin -orbit coupling (SOC), with the broken inversion symmetry alone adequate \nto its emergence .10 Generally, the resulting orbital current is polarized along y direction \nwhen the charge current along x direction, while the orbital current flows toward z \ndirection similar to spin current .11 This orbital current (to distinguish it from spin \ncurrent, denot ing as oy) can generate an in -plane antidamping -like torque (orbital torque , \nOT) 𝜏DL that contributes to highly efficient magne tization switching of thin films \nsimilar to spin torque .12 Besides, the out -of-plane antidamping -like torque possess \npotential applications, in which the generation mechanism of this torque is worthing \nstudying for the development of orbital engineering . \nRecently, the out -of-plane antidamping -like torque is widely reported in spin-\nrelated systems that generated by σz-polarized spin current , which prevails in specific \nsystem with lower crystalline or magnetic symmetry, such as anti-ferromagnetic (AFM) \nmaterials Mn 3GaN13, two-dimensional semimetals WTe 214 as well as \nferromagnet/ferroelectric multiferroic (FM/ FE) heterostructures interface .15 Meanwhile, \nthe confirmation of orbital current becomes more popular issue, where more and more \nsystems are confirmed that exist orbital torque, such as Ta/Ni1, Ru/Ni16, Nb/Co17, etc. \nIn th ese systems, the orbital current induc ed by OHE and the OT is proved to exist in \nthe form of in -plane damping -like torque, which is generated by oy. Besides, the orbital \ncurrent has been theoretically and experimentally verified in naturally oxidized Cu (CuOx) film induced by ORE18-19, which the form of orbital current is still oy. However, \nthe out -of-plane polarized orbital current ( oz) is rarely reported despite its rich orbital \nphysics in orbitronics. \nHerein, the oz-polarized orbital current and out -of-plane antidamping -like torque \nare observed in FM/CuO x bilayers via spin -torque ferromagnetic resonance (ST -FMR). \nThe orbital current generated by ORE in the FM/CuO x interface as shown in Fig. 1a, \nthe induced orbital current are injected into FM and next converted to spin through SOC \nof FM. From the magnetic field dependence of ST -FMR spectra, excepting for oy-\npolarized orbital current, we further observe out -of-plane antidamping -like torque \ngenerated by oz-polarized orbital current. After fitting the ST-FMR antisymmetric \nresonance component Va vs in-plane magnetic -field angle φ and symmetric component \nVs vs φ, respectively, we conclude the magnitude of various OT and the equivalent OT \nefficiency. Besides that, we also characterize the FM species dependence of OT \nefficiency, which result coincides with orbital -related theory. \n2. Experimental \nThe FM/Cu thin films were grown on thermally oxidized Si substrates via \nmagnetron sputtering at room temperature. For clarif ying the physical mechanism of \nthe OT, two different types of bilayers were prepared: Co(8 nm)/Cu( tCu) and Fe 20Ni80(8 \nnm)/Cu(tCu) (Fe 20Ni80 = Permalloy = Py) . The devices were prepare d by exploiting \nshadow mask prior to the deposition and the dimension is 100 × 600 μm2 rectangular \nstrips. The FM/CuO x was obtained via the FM/Cu bilayers exposed to the laboratory \nambient. \n \nFig. 1 (a) Schematic illustration for the mechanism of orbital torque in FM/CuO x. (b) Schematic of \nthe device geometry and the circuit for ST -FMR measurements. \nTo measure the charge -to-orbital conversion, we use the ST -FMR technique as \nx y\nz(a) (b)shown in Fig. 1b . In the ST -FMR measurement, a radio frequency (RF) microwave \ncurrent was applied in the FM/CuO x interface produc ing alternating torques into the \nFM layer, which excites the FM magnetization to precess and drives the resistance \noscillations owing to the anisotropic magnetoresistance (AMR) of the FM layer when \nthe microwave frequency and the external magnetic field satisfy the ferromagnetic \nresonance (FMR) condition20, \n 𝑓=𝛾\n2𝜋√𝐻(𝐻+4𝜋𝑀𝑒𝑓𝑓). (1) \nHere, 𝑀𝑒𝑓𝑓 is the effective saturation magnetization and 𝛾 is the gyromagnetic ratio. \nWe measure a direct current ( DC) voltage Vdc across the stripe that couples the \noscillating resistance and the RF current. In the ST -FMR measurements, this DC signal \nVdc can be fitted with the equation21: \n 𝑉𝑑𝑐=𝑉𝑎∆𝐻(𝐻−𝐻𝑟)\n(𝐻−𝐻𝑟)2+∆𝐻2+𝑉𝑠∆𝐻2\n(𝐻−𝐻𝑟)2+∆𝐻2, (2) \nwhere ∆H is the resonance linewidth, and Hr is the resonance field. Vs and Va are the \nmagnitude of the symmetric and antisymmetric component, respectively. The out -of-\nplane ( 𝜏⊥) and in -plane ( 𝜏ǁ) torques defined in Fig . 1b could be obtained individually, \nas the antisymmetric ST -FMR resonance component Va and symmetric component Vs \nare proportional to the amplitude of the out -of-plane and in -plane torques. These \ncomponents relate to the two torques as follows13: \n𝑉𝑠=−𝐼𝑅𝐹\n2(𝑑𝑅\n𝑑𝜑)1\n𝛼𝛾(2𝐻𝑟+𝜇0𝑀𝑒𝑓𝑓)𝜏ǁ, (3) \n 𝑉𝑎=−𝐼𝑅𝐹\n2(𝑑𝑅\n𝑑𝜑)√1+𝜇0𝑀𝑒𝑓𝑓/𝐻𝑟\n𝛼𝛾(2𝐻𝑟+𝜇0𝑀𝑒𝑓𝑓)𝜏⊥. (4) \nwhere φ is the in -plane magnetic field angle defined in Fig . 1b. dR/dφ is due to the \nAMR in the FM. IRF is the radio microwave current. 𝜇0 is the permeability in vacuum, \nα is the Gilbert damping coefficient. \n2 Results and discussion \nSimilar to spin-related systems, considering only orbital Hall effect (or the orbital \nRashba effect and Oersted field), the out-of-plane and in -plane torques could be \ngenerally given 𝒎×𝒐𝒚 and 𝒎×(𝒐𝒚×𝒎), respectively .22 Under the circumstances , \nanalogous to systems with 2 -fold rotational symmetry, if m is reversal by applying a negative magnetic field equivalent to rotating the in -plane magnetic field angle φ (in \nregard to x direction) by 180º, Vdc must satisfy same amplitude but opposite sign, that \nis Vdc(H) = -Vdc(-H). Any difference in resonance line shape between Vdc(H) and -Vdc(-\nH) represents the existence of an additional torque. Generally, besides oy-induced OT, \nthe oz-induced OT , with features of more comparable Vs signals (both the magnitude \nand sign are taken into account) and distinct Va signals , which may be considered. \n \nFig. 2 ST-FMR resonance signal for (a) Co/CuO x(3 nm), (b) Co/CuO x(10 nm), (c) Py/CuO x(3 nm) \nand (d) Py/CuO x(10 nm). \nFig. 2a shows the representative ST-FMR resonance signal Vdc of the Co/CuO x(3 \nnm) at frequenc y of 10 GHz with the current flow for the magnetic field angle 𝜑= 45° \n(sweeping from the positive and negative magnetic field) at room temperature. \nAccording to previous analysis, comparing the ST -FMR resonance signals under \npositive and negative magnetic field, Va signals with opposite signs reflects an existence \nof oz-induced OT . Moreover, as presented in Fig. 2b for Co/CuO x(10 nm), Vs signals \nwith same signs but smaller magnitude relative to Va means weak er contribution from \noz-induced OT when the thickness of CuO x reaching to 10 nm. One representative \ncharacteristic of the magnitude of OT is strongly dependent on material variation of a \nFM.2 Thus, we perform similar ST -FMR measurements for Py/CuO x(3 nm) bilayers as \nshown in Fig. 2c, distinct Vs and Va signals in magnitude provides obvious evidence for \nthe existence of oz-induced OT. While as for Py/CuO x(10 nm) bilayers, there is no \nobvious difference between Vdc(H) and -Vdc(-H), which indicates the weak oz-induced \nOT. \n \nFig. 3 In-plane magnetic field angle φ dependence of Vs and Va for (a) Co/CuO x(3 nm), (b) \nCo/CuO x(10 nm), (c) Py/CuO x(3 nm) and (d) Py/CuO x(10 nm). \nTo analyse the torque components and magnitude in FM/ CuO x systems, we \nconsider the full angular dependence of the ST -FMR signal as an external magnetic \nfield. The ST -FMR measurement is conducted for values of the in -plane magnetic -field \nangle φ from 0º to 360º at an excitation frequency of 1 0 GHz . To obtain Vs and Va, the \nresonance spectr a are fitted via Eq. (2). The angular dependence of Vs and Va are \nexhibited in Fig. 3a for Co/CuO x(3 nm) bilayers , which can be fitted by the variation of \nEq. (3) and (4) that is 𝑉𝑠 ∝sin(2𝜑)𝜏ǁ and 𝑉𝑎 ∝sin(2𝜑)𝜏⊥ . Traditionally, \nanalogous to spin-related systems, the current -induced orbital torques follow a cos ( φ) \nbehavior due to the orbital Hall effect and/or Oersted field20,23, which leads to a total \nangular dependence of the form sin (2 φ) cos ( φ) for both Vs and Va. However, it is \nnotable that the angular dependence both Va and Vs for the Co/CuO x(3 nm) bilayers \nneed to be well fitted through adding extra term, that is: \n𝑉s=𝐴ssin(2𝜑)cos(𝜑)+𝐵ssin(2𝜑) (5) \n𝑉a=𝐴asin(2𝜑)cos(𝜑)+𝐵asin(2𝜑) (6) \n0 90 180 270 360-0.4-0.20.00.20.4\nCo-CuOx \ntCuOx = 3 nmVdc (V)\n () Vs Assin2cos+ Bssin2\n \n Va Aasin2cos+ Basin2\n0 90 180 270 360-6-3036\nPy-CuOx \ntCuOx = 3 nmVdc (V)\n () Vs Assin2cos+ Bssin2\n \n Va Aasin2cos+ Basin2\n0 90 180 270 360-46-2302346\nPy-CuOx \ntCuOx = 10 nmVdc (V)\n () Vs Assin2cos+ Bssin2\n \n Va Aasin2cos+ Basin2\n0 90 180 270 360-4-2024\nCo-CuOx \ntCuOx = 10 nmVdc (V)\n () Vs Assin2cos+ Bssin2\n \n Va Aasin2cos+ Basin2(a) (b)\n(c) (d)where As(a) and Bs(a) represent the magnitude of OT τy,AD(FL) and τz,FL(AD) generated by \norbital current that are polarized along y direction ( oy) and z direction ( oz), respectively. \nIt is essential to point that the out -of-plane field -like torque τy,FL is deemed to the \ncontribution from current -induced Oersted field .13 In other words, the magnitude of \ncorresponding OT could be obtained by fitting Va(s) as functions of φ. As shown in Fig. \n3a, better fit through two terms further indicates the existence of oz-induced OT. \nHowever, as for Co/CuO x(10 nm) bilayers in Fig. 3b , small er Vs and a trend of close to \nsin (2 φ) cos ( φ) for Va means weaker oz-induced OT. Moreover, in Py/ CuO x bilayers, \nsimilar behavior has been observed , that is, strong oz-induced OT appears at a CuO x \nthickness (tCuOx) of 3 nm in Fig. 3c , while weaken at 10 nm in Fig. 3d. \n \nFig. 4 (a) 𝜏𝑦,𝐴𝐷\n𝜏𝑦,𝐹𝐿 and (b) 𝜏𝑧,𝐴𝐷\n𝜏𝑦,𝐹𝐿 corresponding to in -plane antidamping -like torque and out -of-plane \nantidamping -like torque generated by orbital polarization along y and z direction for Co/CuO x and \nPy/CuO x bilayers , respectively. \nFurthermore, to obtain the proportion of antidamping -like torque in the generated \ntorques , we employ the OT ratios 𝜏𝑦,𝐴𝐷\n𝜏𝑦,𝐹𝐿 and 𝜏𝑧,𝐴𝐷\n𝜏𝑦,𝐹𝐿 corresponding to in-plane \nantidamping -like torque and out -of-plane antidamping -like torque generated by orbital \npolarization along y and z direction , respectively. We extract the magnitude of OT from \nVa(s) vs φ under different tCuOx for Co/CuO x and Py/CuO x bilayers and conclude the OT \nratios , showing tCuOx dependence of 𝜏𝑦,𝐴𝐷\n𝜏𝑦,𝐹𝐿 in Fig. 4a. when the tCuOx from 2 nm to 3 nm, \nthe oy-induced OT ratios 𝜏𝑦,𝐴𝐷\n𝜏𝑦,𝐹𝐿 increase s dramatically and peak s when tCuOx reaches 3 \nnm, highlighting that the orbital current stems from FM/CuO x interface. For thicker \nsamples ( tCuOx > 4 nm), the system becomes FM/Cu/CuO x, it exhibits a monotonic \n3 6 9 120.000.050.100.150.20\ntCuOx (nm)z,AD/y,FL Co-CuOx\n Py-CuOx\n3 6 9 120.00.61.21.82.4\ntCuOx (nm)y,AD/y,FL Co-CuOx\n Py-CuOx(a) (b)decrease of OT ratios as tCuOx increases, which can be attributed to either the current \nshunting of low resistivity Cu o r the blocking of the orbital current in non -oxidized Cu \nwith filled 3d-shell .18 Moreover, the oz-induced OT ratios (𝜏𝑧,𝐴𝐷\n𝜏𝑦,𝐹𝐿) also shows similar \ntrend as shown in Fig. 4b . It is obvious that there exists out-of-plane antidamping -like \ntorque when tCuOx is 3 nm. We contribute the generation of orbital current to ORE19 and \nspin-vorticity coupling (SVC) .24 Note that the oxygen gradient has a vital effect on OT \nand the modulation of orbital current as the electronic configuration of CuO x alters to \nd8 or d9 not d10. According to above result, we conclude that thinner Cu layer ( tCu < 4 \nnm) is all oxidized to CuO x and adjacent to FM layer, where the OAM is absorbed by \nthe magnetization of FM layer. However, the thicker Cu layer, where the system \nexhibits the structure of FM/Cu -CuO x due to shorter length scale of the oxygen gradient \n(~ 3-5 nm), hence the orbital current cannot efficiently diffuse through the non -oxidized \nCu due to intrinsic quenching of the OAM , which could be ascribed to the negligible d \ncharacter of non-oxidized Cu near the Fermi level .19 \n \nFig. 5 (a) Frequency dependence of linewidth ΔH for different CuO x thicknesses. (b) CuO x thickness \ndependence of damping parameter α. \nTo further identify the existence the antidamping -like torque in the system, we \ncarry out microwave frequency f dependence of ST-FMR measurement for \nPy/CuO x(tCuOx) bilayers to obtain the variety of damping .25 Following, we extract \nresonance linewidth ∆𝐻 under different frequency for each sample, showing the f \ndependence of ΔH in Fig. 5a and fitted by following equation26: ∆𝐻=∆𝐻0+4𝜋𝛼𝑓 /𝛾, \nwhere ∆𝐻0 and α can be quantitatively obtained by linear fitting. Almost -zero \ninhomogeneous linewidth broadening ∆𝐻0 implies small roughness of each sample. \n0 5 10 1511223344\nPy-CuOx(t) 2 nm\n 3 nm\n 5 nm\n 6 nm\n 8 nm\n 10 nm\n 12 nmH\nFrequency (GHz)\n0 3 6 9 124.24.54.85.1\nPy-CuOx (x10-3)\ntCuOx (nm)(a) (b)Moreover, w hen the thickness of CuO x gradually increase, the slope of ΔH vs f decrease , \nwhich means a similar trend of α and further characterize s in Fig. 5b. With the \nincreasing CuO x thickness , we find a monotonic decrease of damping α, which proves \nthe existence of anti -damping25 (decrease of α) and attributes to interfacial OT by ORE \nat Py/CuO x system. We note that similar trend w as observed in Py/Ag/Bi26, Py/Cu/Ta27 \nand Ta/Py25 interfacial Rashba systems , where the decrease of α indicates the generation \nof antidamping -like torque. \n3 Conclusion s \nIn conclusion , the generation of out -of-plane anti damping -like torque is \ninvestigated based on orbital Rashba effect in FM/CuO x bilayers by employing spin -\ntorque ferromagnetic resonance. The magnetic field dependence of resonance signal \nproves the existence of both y- and z-polarized orbital current. The oy- and oz- induced \norbital torque ratios 𝜏𝑦,𝐴𝐷\n𝜏𝑦,𝐹𝐿 and 𝜏𝑧,𝐴𝐷\n𝜏𝑦,𝐹𝐿 all increase dramatically and peak when the \nthickness of CuO x reaches 3 nm and then monotonically decrease for thicker CuO x, \nhighlighting that the orbital current originates in FM/CuO x interface and strong \nantidamping -like torque present in 3 -nm CuO x systems, which further be proved by \nCuO x thickness dependence of damping parameter α. This result provides a physic \nunderstanding of the generation out-of-plane antidamping -like torque by orbital current \nand opens a route to generating efficient OT by orbital engineering. \n \nConflict of Interest \nThe authors declare no conflict of interest. \n \nAcknowledgements \nThis work was supported by the National Natural Science Foundation of China (Grant \nNo. 52271179 ), the Natural Science Foundation of Gansu Province (Grant No. \n21JR7RA472). \n Notes and r eferences \n1 D. Lee, D. Go, H. J Park, W. Jeong, H. W. Ko, D. Yun, D. Jo, S. Lee, G. Go, J. H. Oh, \nK. J. Kim, B. G. Park, B. Ch. Min, H. Ch. Koo, H. W. Lee, O. Lee and K. J. Lee, Nat. \nCommun ., 2021, 12, 6710. \n2 D. Go and H. W. Lee , Phys. Rev. Research , 2020, 2, 013177 . \n3 L. Salemi and P. M. Oppeneer, Phys. Rev. Mater. , 2022, 6, 095001. \n4 Sh. L. Ding, A. Ross, D. W. Go, L. Baldrati, Z. Y . Ren, F. Freimuth, S. Becker, F. \nKammerbauer, J. B. Yang, G. Jakob, Y . Mokrousov, and M. Kläui , Phys. Rev. Lett. , \n2020, 125, 177201. \n5 S. Lee, M. G. Kang, D. Go, D. Kim, J. H. Kang, T. Lee, G. H. Lee, J. Kang, N. J. Lee, \nY . Mokrousov, S. Kim, K. J. Kim, K. J. Lee and B. G. Park, Commun. Phys. , 2021, 4, \n234. \n6 T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hirashima, K. Yamada, and J. Inoue \nPhys. Rev. B , 2008, 77, 165117. \n7 D. Jo, D. Go, and H. W. Lee , Phys. Rev. B , 2018, 98, 214405. \n8 J. H. Park, Ch. H. Kim, H. W. Lee, and J. H. Han, Phys. Rev. B , 2013, 87, 041301(R). \n9 S. R. Park, Ch. H. Kim, J.Yu, J. H. Han, and Ch. Kim, Phys. Rev. Lett. , 2011, 107, \n156803. \n10 J. Kim, D. Go, H. Tsai, D. Jo, K. Kondou, H. W. Lee, and Y . Otani, Phys. Rev. B , \n2021, 103, L020407. \n11 J. Sinova , S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Rev. Mod. \nPhys. , 2015, 87, 1213. \n12 A. Manchon, J. Železný, I.M. Miron, T. Jungwirth, J. Sinova, A. Thiaville, K. \nGarello, and P. Gambardella , Rev. Mod. Phys. , 2019, 91, 035004. \n13 T. Nan, C. X. Quintela, J. Irwin, G. Gurung, D. F. Shao, J. Gibbons, N. Campbell, \nK. Song, S. -Y . Choi, L. Guo, R. D. Johnson, P. Manuel, R. V . Chopdekar, I. Hallsteinsen, \nT. Tybell, P. J. Ryan, J. W. Kim, Y . Choi, P. G. Radaelli, D. C. Ralph, E. Y . Tsymba l, M. \nS. Rzchowski, and C. B. Eom, Nat. Commun. , 2020, 11, 4671. \n14 D. MacNeill, G. M. Stiehl, M. H. D. Guimaraes, R. A. Buhrman, J. Park, and D. C. Ralph, Nat. Phys. , 2017, 13, 300 -305. \n15 F. Liu, Y . Jin, Y . Zhao, Ch. Jia, and Ch. Jiang , Appl. Phys. Lett., 2022, 121, 022404 . \n16 A. Bose, F. Kammerbauer, R. Gupta, D. Go, Y . Mokrousov, G. Jakob, and M. Kläui , \nPhys. Rev. B , 2023, 107, 134423 . \n17 F. Liu, B. Liang, J. Xu, Ch. Jia, and Ch. Jiang , Phys. Rev. B , 2023, 107, 054404. \n18 Sh. Ding, Zh . Liang, D. Go, Ch. Yun, M. Xue, Zh. Liu, S. Becker, W. Yang, H. Du, \nCh. Wang, Y . Yang, G. Jakob, M. Kläui, Y . Mokrousov, and J. Yang, Phys. Rev. Lett. , \n2022, 128, 067201. \n19 D. Go, D. Jo, T. Gao, K. Ando, S. Blügel, H. W. Lee, and Y . Mokrousov , Phys. Rev. \nB, 2021, 103, L121113. \n20 L. Q. Liu, T. Moriyama, D. C. Ralph, and R. A. Buhrman, Phys. Rev. Lett. , 2011, \n106, 03660. \n21 W. L. Yang, J. W. Wei, C. H. Wan, Y . W. Xing, Z. R. Yan, X. Wang, C. Fang, C. Y . \nGuo, G. Q. Yu, and X. F. Han , Phys. Rev. B , 2020, 101, 064412. \n22 Y . Wang, R. Ramaswamy, and H. Yang, J. Phys. D: Appl. Phys. , 2018, 51, 273002. \n23 A. R. Mellnik, J. S. Lee, A. Richardella, J. L. Grab, P. J. Mintun, M. H. Fischer, A. \nVaezi , A. Manchon, E. A. Kim, N. Samarth, D. C. Ralph, Nature , 2014, 511, 449 -451. \n24 Genki Okano, Mamoru Matsuo, Yuichi Ohnuma, Sadamichi Maekawa, and Yukio \nNozaki , Phys. Rev. Lett. , 2019, 122, 217701. \n25 N. Behera, S. Chaudhary, and D. Pandya, Sci. Rep. , 2016, 6, 19488. \n26 M. B. Jungfleisch, W. Zhang, J. Sklenar, W. Jiang, J. E. Pearson, J. B. Ketterson, and \nA. Hoffmann, Phys. Rev. B , 2016, 93, 224419. \n27 F. Liu, C. Zhou, R. Tang, G. Chai, and C. Jiang, J. Magn. Magn. Mater. , 2021, 540, \n168462 . " }, { "title": "0710.2826v2.Ferromagnetic_resonance_study_of_polycrystalline_Fe__1_x_V_x_alloy_thin_films.pdf", "content": "arXiv:0710.2826v2 [cond-mat.mes-hall] 17 Oct 2007Ferromagnetic resonance study of polycrystalline Fe 1−xVxalloy thin films\nJ-M. L. Beaujour, A. D. Kent\nDepartment of Physics, New York University, 4 Washington Pl ace, New York, NY 10003, USA\nJ. Z. Sun\nIBM T. J. Watson Research Center, Yorktown Heights, NY 10598 , USA\n(Dated: November 18, 2018)\nFerromagnetic resonance has been used to study the magnetic properties and magnetization dy-\nnamics of polycrystalline Fe 1−xVxalloy films with 0 ≤x <0.7. Films were produced by co-\nsputtering from separate Fe and V targets, leading to a compo sition gradient across a Si substrate.\nFMR studies were conducted at room temperature with a broadb and coplanar waveguide at fre-\nquencies up to 50 GHz using the flip-chip method. The effective demagnetization field 4 πMeffand\nthe Gilbert damping parameter αhave been determined as a function of V concentration. The\nresults are compared to those of epitaxial FeV films.\nI. INTRODUCTION\nA decade ago, it was predicted that a spin polar-\nized current from a relatively thick ferromagnet (FM)\ncould be used to switch the magnetization of a thin FM\n[1]. Since then, this effect, known as spin-transfer, has\nbeen demonstrated in spin-valves [2] and magnetic tun-\nnel junctions [3]. In a macrospin model with collinear\nlayer magnetizations, there is a threshold current den-\nsityJcfor an instability necessary for current-induced\nmagnetization switching of the thin FM layer [1, 4]:\nJc=2eαMstf(Hk+2πMs)\n/planckover2pi1η, (1)\nwhereαisthedampingconstant. tfandMsarethethick-\nness and the magnetization density of the free layer, re-\nspectively. Hkis the in-plane uniaxial anisotropy field. η\nis the currentspin-polarization. In orderfor spin-transfer\nto be used in high density memory devices Jcmust be re-\nduced. From Eq. 1 it is seen that this can be achieved by\nemploying materials with low Msandαin spin-transfer\ndevicesor, equivalently materialswith lowGilbert damp-\ning coefficients, G = αMs(gµB//planckover2pi1).\nVery recently, an experimental study of epitaxial FeV\nalloy thin films demonstrated a record low Gilbert damp-\ning coefficient [5]. This material is therefore of interest\nfor spin transfer devices. However, such devices are gen-\nerally composed of polycrystalline layers. Therefore it is\nof interest to examine polycrystalline FeV films to assess\ntheir characteristics and device potential.\nIn this paper, we present a FMR study of thin poly-\ncrystalline Fe 1−xVxalloy films with 0 ≤x <0.7 grown\nby co-sputtering. The FeV layers were embedded be-\ntween two Ta |Cu layers, resulting in the layer structure\n||5 Ta|10 Cu|FeV|5 Cu|10 Ta||, where the numbers are\nlayer thickness in nm. FeV polycrystalline films were\nprepared by dc magnetron sputtering at room tempera-\nture from two separate sources, oriented at a 45oangle\n(Fig. 1a). The substrate, cut from a Silicon wafer with\n100 nm thermal oxide, was 64 mm long and about 5 mm\nwide. The Fe and V deposition rates were found to vary/s49 /s50 /s51 /s52/s45/s50/s45/s49/s48/s40/s98/s41\n/s32/s86/s32/s116/s97/s114/s103/s101/s116\n/s52/s53/s111\n/s32\n/s32/s115/s117/s98/s115/s116/s114/s97/s116/s101/s70/s101/s32/s116/s97/s114/s103/s101/s116/s40/s97/s41\n/s72\n/s114/s101/s115/s65/s98/s115/s111/s114/s112/s116/s105/s111/s110/s32/s108/s105/s110/s101/s32/s32/s40/s97/s117/s41\n/s72\n/s97/s112/s112/s32/s32/s40/s107/s79/s101/s41/s49/s52/s32/s71/s72/s122\n/s32/s32/s32\n/s32/s120 /s61/s48/s46/s51/s55/s120 /s61/s48/s46/s53/s50\n/s120 /s61/s48/s46/s49/s57/s72\n/s115/s117/s98/s115/s116/s114/s97/s116/s101/s32/s104/s111/s108/s100/s101/s114\n/s97/s120/s105/s115\nFIG. 1: a) The co-sputtering setup. b) Typical absorp-\ntion curves at 14 GHz for a selection of ||Ta|Cu|7.5 nm\nFe1−xVx|Cu|Ta||films, with x=0.19, 0.37 and 0.52. The res-\nonance field Hresand the linewidth ∆ Hare indicated.\nlinearly across the wafer. The Fe and V rates were then\nadjusted to produce a film in which xvaries from 0.37\nto 0.66 across the long axis of the wafer. The base pres-\nsure in the UHV chamber was 5 ×10−8Torr and the\nAr pressure was set to 3.5 mTorr. The FeV was 7.5 nm\nin thickness, varying by less than 0.3 % across the sub-\nstrate. An Fe 1−xVxfilm 3 nm thick was also fabricated.\nTo produce films with x <0.30 the rate of the V source\nwas decreased. Finally, pure Fe films with a thickness\ngradient ranging from 7 nm to 13.3 nm were deposited.\nTheFMRmeasurementswerecarriedoutatroomtem-\nperature using a coplanar wave-guide (CPW) and the\nflip-chip method. Details of the experimental setup and\nofthe CPWstructuralcharacteristicscan be found in [6].\nAdc magnetic field, up to 10 kOe, wasapplied in the film\nplane, perpendiculartotheacmagneticfield. Absorption\nlines at frequencies from 2 to 50 GHz were measured by\nmonitoring the relative change in the transmitted signal\nas a function of the applied magnetic field.2\n/s48/s46/s48 /s48/s46/s51 /s48/s46/s54/s48/s53/s49/s48/s49/s53/s50/s48\n/s48/s46/s48 /s48/s46/s51 /s48/s46/s54/s50/s46/s48/s53/s50/s46/s49/s48/s50/s46/s49/s53/s49/s50/s51\n/s120\n/s32/s52 /s77\n/s101/s102/s102/s32/s32/s40/s107/s71/s41/s55/s46/s53/s32/s110/s109/s32/s70/s101\n/s49/s45 /s120/s86\n/s120\n/s51/s32/s110/s109/s32/s70/s101\n/s48/s46/s54/s51/s86\n/s48/s46/s51/s55\n/s49/s50/s46/s57/s32/s110/s109/s32/s70/s101\n/s120/s32\n/s32/s103/s45/s102/s97/s99/s116/s111/s114/s40/s98/s41/s40/s99/s41/s32\n/s32\n/s32/s72\n/s114/s101/s115/s32/s32/s40/s107/s79/s101/s41\n/s49/s49/s32/s71/s72/s122/s40/s97/s41\nFIG. 2: a) The resonance field at 11 GHz versus xand b) the\neffective demagnetization field versus x. The solid line is a\nguide to the eye. c) The Land´ e gfactor as a function of x.\nThe dotted line shows the g-factor value of bulk Fe.\nII. RESULTS\nTypical absorption lines at 14 GHz of selected FeV al-\nloyfilmsareshowninFig. 1b. Thelinesarelorentzianfor\nmost frequencies. At a fixed frequency, the FMR absorp-\ntion decreaseswith increasingV content. The FMR peak\nof a film 7.5 nm thick with x= 0.66 is about 100 times\nsmallerthanthatofapureFeofthesamethickness. This\nis accompaniedby a shift of Hrestowardshigher field val-\nues (Fig. 2a). The effective demagnetization field 4 πMeff\nand the Land´ eg-factorgwere determined by fitting the\nfrequency dependence of the resonance field Hresto the\nKittel formula [7]:\nf2=/parenleftBiggµB\nh/parenrightBig2\nHres(Hres+4πMeff),(2)\nwhere the effective demagnetization field is:\n4πMeff= 4πMs−H⊥. (3)\nNote that in the absence of a perpendicular anisotropy\nfieldH⊥, the effective field would be directly related to\nMs. The dependence of 4 πMeffon V concentration is\nshown in Fig. 2c. As xincreases the effective demagneti-\nzation field decreases dramatically, going from about 16\nkG forx= 0 to 1.1 kG for x= 0.66. Note that the effec-\ntive demagnetization field of the 7.5 nm Fe film is about\n25 % lower than that of bulk Fe (21.5 kG). The 12.9\nnm Fe film exhibits a larger 4 πMeff, which is, however,\nstill lower than 4 πMsof the bulk material. Similarly,\nthe 4πMeffof an Fe 0.63V0.37film is thickness dependent:\ndecreasing with decreasing layer thickness.\nThe Land´ eg-factor increases monotonically with in-\ncreasing V concentration (Fig. 2b). The minimum g-\nfactor is measured for the Fe film: g= 2.11±0.01,\nwhich is slightly larger than the value of bulk material\n(g= 2.09). Note that gof a Fe film 12.9 nm thick is\nequal to that of Fe bulk. However, the g-factor of the/s48 /s50/s48 /s52/s48/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\n/s48/s46/s48/s48/s46/s51/s48/s46/s54\n/s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54/s48/s46/s53/s49/s46/s48/s49/s46/s53 /s40/s99/s41/s120 /s32/s61/s32/s48/s46/s53/s50/s120/s32 /s61/s32/s48/s46/s52/s51/s120/s32 /s61/s32/s48/s40/s98/s41/s32/s40/s49/s48/s45/s50\n/s41/s76/s105/s110/s101/s119/s105/s100/s116/s104/s32/s32/s40/s107/s79/s101/s41/s32\n/s32/s72 /s32/s32/s40/s107/s79/s101/s41\n/s72\n/s48/s72\n/s48/s72\n/s49/s52/s32/s32\n/s32/s40/s97/s41\n/s120/s32 /s32/s102/s32/s32/s40/s71/s72/s122/s41\nFIG. 3: a)Frequency dependence of the linewidth for 7.5 nm\nFe1−xVxalloy film with x=0, 0.43 and 0.52. The solid lines\nare the best linear fit of the experimental data. b) ∆ H14, the\nlinewidth at 14 GHz , and ∆ H0are shown as a function of x.\nc) The magnetic damping parameter versus V concentration.\nFe0.63V0.37, does not appear to be thickness dependent:\nthe 3 nm Fe 0.63V0.37layer has about the same gvalue\nthan the 7.5 nm Fe 0.63V0.37layer.\nThe half-power linewidth, ∆ H, was studied as a func-\ntion of the frequency and of the V concentration. Fig.\n3b shows the dependence of the FMR linewidth on xat\n14 GHz. The general trend is that ∆ Hincreases with\nx. However, there are two regimes. For x >0.4, the\nlinewidth depends strongly on x, increasing by a factor\n5 whenxis increased from 0.4 to 0.66. The dependence\nof the linewidth on xis more moderate for the films with\nx <0.4: it increases by about 30 %. For all samples,\nthe linewidth scales linearly with the frequency. A least\nsquare fit of ∆ H(f) gives ∆ H0, the intercept at zero\nfrequency, and the Gilbert damping parameter αwhich\nis proportional to the slope: d∆H/df= (2h/gµB)α[8].\n∆H0is typically associated with an extrinsic contribu-\ntion to the linewidth and related to magnetic and struc-\ntural inhomogeneities in the layer. For two samples with\nthe highest Vanadium concentration, x= 0.60, 0.66, the\nlinewith is dominated by inhomogeneous broadening and\nit wasnot possible to extract α. Asxincreases, ∆ H0and\nαincreases. The damping parameter and ∆ H0remain\npractically unchanged for x≤0.4 and when x >0.4,\nboth the intercept and the slope of ∆ Hversusfincrease\nrapidly.\nIII. DISCUSSION\nSeveral factors can contribute to the dependence of\n4πMeffon the V concentration. First, the decrease of\nthe effective demagnetizationfield canbe associatedwith\nthe reduction of the alloy magnetization density Mssince\nthe Fe content is reduced. In addition, a neutron scat-\ntering study showed that V acquires a magnetic moment3\nantiparallel to the Fe, and that the Fe atom moment de-\ncreases with increasing V concentration [9]. The Curie\ntemperature of Fe 1−xVxdepends on x. In fact, Tcfor\nx=0.65 is near room temperature [10]. It is important to\nmention that a factor that can further decrease 4 πMeff\nis an out-of-plane uniaxial anisotropy field H⊥(Eq. 3).\nIn thin films, the perpendicular anisotropy field is com-\nmonly expressed as H⊥= 2K⊥/(Mst), where K⊥>0\nis the anisotropy constant and tthe ferromagnetic film\nthickness [11]. In this simple picture, it is assumed that\nK⊥is nearly constant over the thickness range of our\nfilms. This anisotropy can be associated with strain\ndue to the lattice mismatch between the FeV alloy and\nthe adjacent Cu layers and/or with an interface contri-\nbution to the magnetic anisotropy. For Fe films with\nt= 7.5 and 12.9 nm, a linear fit of 4 πMeffversus 1/t\ngives 4πMs= 20.2 kG and K⊥= 2.5 erg/cm2. The\nvalue extracted for 4 πMsis in the range of the value of\nFebulk. AsimilaranalysisconductedonFe 0.63V0.37films\nof thickness t=3 and 7.5 nm gives 4 πMs= 12.2 kG and\nK⊥= 0.1 erg/cm2. The result suggests that the surface\nanisotropy constant decreases with increasing x.\nIV. SUMMARY\nThe effective demagnetization field of the polycrys-\ntalline Fe 1−xVxalloy films decreases with increasing xandalmostvanishesfor x≈0.7. AFMRstudyonepitax-\nial films haveshown a similar xdependence of 4 πMeff[5].\nUsing the value of Mscalculated in the analysis above,\nwe estimate the Gilbert damping constant of a 7.5 nm Fe\nlayer and 7.5 nm Fe 0.63V0.37alloy film to be G Fe= 239\nMHz and G FeV= 145 MHz respectively. The decrease of\nthe effective demagnetization field of Fe 1−xVxwith in-\ncreasing xis accompanied by a decrease of the Gilbert\ndamping constant. A similar xdependence of G was\nobserved in epitaxial films [5]. The authors explained\nthe decrease of G by the reduced influence of spin-orbit\ncoupling in lighter ferromagnets. Note that the Gilbert\ndampingofourfilmsislargerthanwhatwasfoundforthe\nepitaxial films (G=57 MHz for epitaxial Fe 8 nm thick).\nWe note that the Fe 0.63V0.37alloy film, which has\n4πMsapproximatly the same as that of Permalloy, has a\nmagnetic damping constant of the same order than that\nof Py layer in a similar layer structure [12]. Hence, with\ntheir low Msandα, polycrystalline FeV alloy films are\npromising materials to be integrated in spin-tranfer de-\nvices.\n[1] J. C. Slonczewski, J. Magn. Magn. Mater. 159(1-2), L1\n(1996) ; L. Berger, Phys. Rev. B 54(12), 9353 (1996).\n[2] see, for example, J. A. Katine et al., Phys. Rev. Lett. 84,\n3149 (2000) ; B. Oezyilmaz et al., Phys. Rev. Lett. 91,\n067203 (2003).\n[3] see, for example, G. D. Fuchs et al., J. Appl. Phys. 85\n(7), 1205 (2004).\n[4] J. Z. Sun, Phys. Rev. B 62(1), 570 (2000).\n[5] C. Scheck et al., Phys. Rev. Lett. 98, 117601 (2007).\n[6] J-M. L. Beaujour et al., Europhys. J. B, DOI:\n10.1140/epjb/e2007-00071-1 (2007).\n[7] C. Kittel in Introduction to Solid State Physics, Ed. 7,\np.505.[8] see, for example, D. L. Mills and S. M. Rezende in Spin\nDynamics in Confined Magnetic Structures II (Eds. B.\nHillebrands andK.Ounadjela), pp.27-58, (Springer, Hei-\ndelberg 2002).\n[9] I. Mirebeau, G. Parette, and J. W. Cable. J. Phys. F:\nMet. Phys. 17, 191 (1987).\n[10] Y. Kakehashi, Phys. Rev. B 32(5), 3035 (1985).\n[11] Y. K. Kim and T. J. Silva, Appl. Phys. Lett. 68, 2885\n(1996).\n[12] S. Mizukami et al., J. Magn. Magn. Mater. 239, 42\n(2002)." }, { "title": "0712.1403v2.Ferromagnetic_Quantum_Critical_Fluctuations_and_Anomalous_Coexistence_of_Ferromagnetism_and_Superconductivity_in_UCoGe_Revealed_by_Co_NMR_and_NQR_Studies.pdf", "content": "arXiv:0712.1403v2 [cond-mat.supr-con] 31 Dec 2007Typeset with jpsj2.cls Full Paper\nFerromagnetic Quantum Critical Fluctuations and Anomalou s Coexistence of\nFerromagnetism and Superconductivity in UCoGe Revealed by Co-NMR and NQR\nStudies\nTetsuya Ohta,1∗YusukeNakai,1Yoshihiko Ihara,1KenjiIshida,1†Kazuhiko Deguchi,2Noriaki K. Sato2\nand Isamu Satoh3\n1Department of Physics, Graduate School of Science, Kyoto Un iversity, Kyoto 606-8502, Japan.\n2Department of Physics, Graduate School of Science, Nagoya U niversity, Nagoya 464-8602, Japan.\n3Institute for Materials Research, Tohoku University, Send ai 980-8577, Japan.\nCo nuclear magnetic resonance (NMR) and nuclear quadrupole resonance (NQR) studies\nwere carried out for the recently discovered UCoGe, in which the ferromagnetic and supercon-\nducting(SC) transitions are reported tooccur at TCurie∼3 K andTS∼0.8 K (N. T. Huy et al.,\nPhys. Rev. Lett. 99(2007) 067006), in order to investigate the coexistence of f erromagnetism\nand superconductivity as well as the normal-state and SC pro perties from a microscopic point\nof view. From the nuclear spin-lattice relaxation rate 1 /T1and Knight-shift measurements, we\nconfirm that ferromagnetic fluctuations that possess a quant um critical character are present\naboveTCurieand also the occurrence of a ferromagnetic transition at 2.5 K in our polycrys-\ntalline sample. The magnetic fluctuations in the normal stat e show that UCoGe is an itinerant\nferromagnet similar to ZrZn 2and YCo 2. The onset SC transition is identified at TS∼0.7 K,\nbelow which 1 /T1arising from 30% of the volume fraction starts to decrease du e to the opening\nof the SC gap. This component of 1 /T1, which follows a T3dependence in the temperature\nrange 0.3−0.1 K, coexists with the magnetic components of 1 /T1showing a√\nTdependence\nbelowTS. From the NQR measurements in the SC state, we suggest that th e self-induced\nvortex state is realized in UCoGe.\nKEYWORDS: UCoGe, ferromagnetic superconductor, Co-NMR an d NQR, coexistence\nSince the discovery of superconductivity in ferromag-\nnetic compounds under high pressure,1,2the concept of\nthe interplay between magnetism and superconductiv-\nity was changed, because ferromagnetism and supercon-\nductivity are considered to be mutually exclusive. The\npressure studies in UGe 2have shown that the super-\nconducting (SC) transition temperature TSis higher in\nthe pressure region where the ordered moments are en-\nhanced,and that superconductivityis not observedwhen\nthe ferromagnetism disappears in a further higher pres-\nsure region.3These results suggest that ferromagnetism\nseems to enhance the superconductivity in UGe 2. In ad-\ndition, ambient-pressure ferromagnetic superconductiv-\nity was discovered in URhGe, where TCurie= 9.5 K and\nTS∼0.25 K were reported.4The relation between the\nferromagnetismand superconductivity observedin the U\ncompounds is one of the attractive topics for the com-\nmunity studying strongly correlated electron systems.\nQuite recently, new ambient-pressure ferromagnetic\nsuperconductivity with TCurie= 3 K and TS∼0.8 K\nwas found in UCoGe.5One of the advantages of UCoGe\nisthatitincludes“NMR-active”nucleusCo.IntheNMR\nstudies of ferromagnetic superconductors, Ge in UGe 2is\nreplaced by73Ge with a nuclear spin and73Ge nuclear\nquadrupole resonance (NQR) studies have been carried\nout by Kotegawa et al.6and Harada et al.7,8They have\nreported the SC properties and the relation between fer-\nromagnetism and superconductivity from the73Ge-NQR\n∗E-mail address: t-ohta@scphys.kyoto-u.ac.jp\n†E-mail address: kishida@scphys.kyoto-u.ac.jp0.0 0.2 0.4 0.6 0.8 1.0 Tc mid Tc onsetpolycrystalline sample δ χac ∝ - δ f (arb. units) UCoGe \n \nT (K) 9 kG \n 7 kG \n 6 kG \n 5 kG \n 4 kG \n 3 kG \n 2 kG \n 1 kG \n 0 kG \nFig. 1. (Color online) Temperature dependence of δχacunder\nvarious magnetic fields below 9 kOe.\nresults. In this paper, we show the detailed Co-NMR and\nNQR results of UCoGe, which are compared with the\n73Ge-NQR results of high-pressure UGe 2.\nPolycrystalline UCoGe was prepared by the arc melt-\ning method. The details will be reported in a separate\npaper.9Figure 1 shows the temperature dependence of\nχacin various magnetic fields, which was obtained by\nmeasuring the self-inductance of an NMR coil with pow-\n12 J. Phys. Soc. Jpn. Full Paper Author Name\n10 15 20 25 59 K = 0 \n1.5 K 4.2 K 6 K 8 K 10 K 15 K 20 K 40 K 60 K 100 K 150 K 63 Cu 65 Cu \nf = 18.2 MHz 59 Co-NMR Intensity (arb. units) UCoGe \nMagnetic Field (kOe) \nFig. 2. (Color online) Field-sweep59Co-NMR spectra obtained\nat 18.2 MHz at various temperatures. The sharp signals shown\nby63Cu and65Cu are the NMR signals arising from the NMR\ncoil.59K= 0 shows the field where the Knight shift of59Co is\nzero. The dashed curve traces the NMR-signal peak arising fr om\nthe 1/2↔ −1/2 transition.\ndered UCoGe. A clear Meissner signal was observed in\nzero and magnetic fields, indicating that superconduc-\ntivity persists up to 9 kOe. In zero field, the onset and\nmidpoint of the SC transition are estimated to be 0.7\nand 0.55 K, respectively. The Meissner behavior in zero\nfield in our sample is consistent with that in a previous\nreport.5\nFigure2showsthe59Co-NMRspectraofourpowdered\nsample obtained by sweeping an external field at various\ntemperatures. A symmetric NMR spectrum with several\nsatellite peaks at higher temperatures, indicative of the\npresence of the electric field gradient (EFG) at the Co\nsite, becomes anisotropic with decreasing temperature.\nThe shift of the gravity of the broad NMR spectra orig-\ninates from the isotropic component of the Knight shift.\nThe NMR peak arising from the 1 /2↔ −1/2 transition\nshifts significantlywith respecttotemperature. Theshift\nof this peak is traced by the dashed curve in Fig. 2.\nFigure 3 shows the temperature dependence of the\nKnight shift59Kevaluated from the peak arising from\nthe 1/2↔ −1/2 transition, along with the magnetiza-\ntionMdivided by Hat 15 kOe. The temperature de-\npendence of59Kis well scaled with M/H. With decreas-\ningT,59Kbecomes so large that it increases up to 35%\nat low temperatures. It is confirmed that the Curie be-\nhavior in M/His microscopically intrinsic. It should be\nnoted that the59Kand Co-NMR spectra do not change\nthrough TCurie∼3 K, showing that the ferromagnetic\ntransition is smeared out in the magnetic fields. This is\ncharacteristic of the ferromagnetic transition. From the\nlinear relationship between59KandM/H, the hyper-1 10 100 010 20 30 40 50 60 \nf = 18.2 MHz \n 59 KUCoGe 59 K (%) \nT (K) 0.00 0.01 0.02 0.03 0.04 \n M/H in 15 kOe \n M/H (emu/mol) 0.00 0.01 0.02 010 20 30 40 59 K (%) \nM/H (emu/mol) \nFig. 3. Temperature dependence of the59Co Knight shift eval-\nuated from the 1 /2↔ −1/2 transition of the spectra shown in\nFig. 2. The temperature dependence of M/Hmeasured in 15\nkOe is shown in the same figure. The inset is the59K−M/H\nplot.\n4 5 8 9(c) \n95 mK νQ= 2.85 MHz \nη = 0.52 \nFrequency (MHz) (b) \n1.5 K 59 Co-NQR Intensity (arb. units) UCoGe \nν\u0001 ν\u0002ν\u0003(a) \n4.2 K \nFig. 4. Co-NQR spectra at 4.2 K (above TCurie: (a)), 1.5 K (be-\nlowTCurie: (b)), and 95 mK (the lowest temperature in the mea-\nsurements: (c)). The NQR spectrum at95 mKcan be understood\nby the superposition of the two NQR spectra arising from the\nregions with and without the internal field. (Refer to the tex t).\nfine coupling constant Hhfat the Co site and the or-\nbital shift Korbare estimated to be 79 ±1 kOe/µBand\n−0.96±0.22%, respectively. From the positive value of\nHhfand the predominance of the isotropic term in the\nKnight shift, it is considered that the core polarization\neffect by Co-3 delectrons is not dominant but the trans-\nferred contribution from U-5 felectrons by way of Co-4 s\norbitsisdominantattheCo-nuclearsite.TheCoKnight-\nshift results indicate that U-5 felectrons play a crucial\nrole for the ferromagnetism in UCoGe.\nThree Co-NQR signals were observed at the peaks of\n4.8, 5.2, and 8.3 MHz at 4.2 K, as shown in Fig. 4. From\nthese signals, we evaluated the NQR parameters νzz,J. Phys. Soc. Jpn. Full Paper Author Name 3\u0000 \u0004 \u0005 \u0006 \u0007 \b \t \n \u000b\n\f \r\n\u000e\n\u000f \u0010\n\u0011\n\u0012 \u0013\n\u0014\n\u0015 \u0016\n\u0017\u0018 \u0019\n\u001a\n\u001b \u001c\n\u001d\n\u001e \u001f\n \n! \"\n#$ % & ' ( ) * + , - .\n/ 0 1\n2\n345 6 7 8\n9 : ; < =>?@\nA\nBC\nDE\nFG H I J\nKL\nM N O P QR S\nT U VWXY\nZ\n[\\\n]^\n_\n` a b c de fg h i j kl m\nn\no p q r\ns t u v\nw x y\nz{ | } ~ \n \n \n \n¡¢ ∞ \n£−\n¤¥¦§¨©ª« ∞ \n¬ ® ¯ ° ± ² ³ ´ µ ¶\n· ¸ ¹º » ¼ ½ ¾ ¿ ÀÁ  ÃÄ Å Æ Ç È É ÊË Ì Í Î Ï Ð Ñ Ò Ó Ô\nFig. 5. (Color online) The main graph is the temperature depe n-\ndence of 1 /T1measured at the Co-NQR signal of 8.3 MHz. The\nvalues of 1 /T1above (below) 1.5 K are shown by open (gray and\nblack) circles with respect to the right (left) axis due to th e dif-\nference in the 1 /T1values by the measurement conditions (refer\nto the text). The scale of the left-hand-sideaxis is 0.58 tha t of the\nright-hand-side axis. The insets are the recoveries of the n uclear\nmagnetization after a saturation pulse. The dotted curves i n the\ninsets are the fitting curves for evaluating 1 /T1. The recoveries\naboveTSare fitted by a single component of 1 /T1, as shown in\nthe insets of (a) and (b). Two components of 1 /T1are observed\nbelowTS, as shown in the inset of (c). The longer (black circles)\nand shorter (gray circles) components of 1 /T1are plotted in the\nmain figure. The dotted curve below TSrepresents the temper-\nature dependence of 1 /T1calculated using the 3-D polar model\nwith 2∆ 0/kBTS= 4.\nwhich is the NQR frequency along the principal axis of\nthe EFG, to be 2.85 MHz and an asymmetric parameter\nηto be 0.52. The large ηvalue shows that the symme-\ntry of the Co site is lower than an axial symmetry. The\nCo-NQR spectra affected by the internal fields at the Co\nsite were observed at low temperatures. The spectrum\nchange by temperature is discussed below.\nThe nuclear spin-lattice relaxation rate 1/ T1of Co in\nzero magnetic field was measured at the NQR signal ob-\nserved at 8.3 MHz denoted as ν3in Fig. 4. Figure 5\nshows the temperature dependence of 1 /T1of Co over\nthe wide temperature range between 95 mK and 150 K.\nIt was found that the values of 1 /T1below 10 K depend\non the NQR pulse conditions. This is considered to be\ndue to the development of the short-range ferromagnetic\ncorrelations. The details of this will be reported else-\nwhere.10Thus, the values of 1 /T1above and below 1.5\nK are shown by different axes because the measurement\nconditions change around 1.5 K. The insets show the re-\ncovery curves m(t) of the nuclear magnetization M(t)1 10 100 050 100 150 200 250 \n0.00 0.05 0.10 0.15 \n0.00 0.05 0.10 0100 200 \nCo-NQR \nf = 8.3 MHz UCoGe \nχ (emu / mol) TCurie \nT (K )1/T 1T \nÕs\nÖ1 K \n×1\nØ\n1/T 1T \nÙs-1 K -1 \nÚ\nχ Ûemu/ molÜ\nFig. 6. (Color online) Temperature dependence of 1 /T1Tin zero\nfield and χbulkmeasured in 1 kOe. The dashed curve shows a\nT−4/3dependence, which is expected at the FM critical point.16\nThe inset is the plot of 1 /T1Tagainstχbulk.\nat a time tafter a saturation pulse at three character-\nistic temperatures. Above TS, the recovery curves were\nalmost consistently determined with a single component\nof 1/T1, as observed in the insets (a) and (b).\nFirst of all, we discuss the magnetic properties in the\nnormal state on the basis of the 1 /T1and59Kresults.\nIn general, 1 /T1divided by temperature 1 /T1Tis related\nto the low-energypart ofthe q-dependent dynamical sus-\nceptibility in compounds. The spin-fluctuation character\nin compounds can be known from the comparison be-\ntween 1/T1Tandχor Knight shift. The main plot of\nFig. 6 shows the temperature dependence of 1 /T1Tin\nzero field, together with that of χbulkin 1 kOe. The good\nlinear relation was found between 1 /T1Tandχbulk, as\nshownintheinset.Thisclearlyprovesthatferromagnetic\n(FM) fluctuations are dominant in UCoGe, since the dy-\nnamical susceptibility revealed by 1 /T1Tis determined\nby theq= 0 susceptibility. The relation1 /T1T∝χbulkis\noften observed in weak and nearly FM compounds,11–13\nand it is interpreted by the SCR theory.14,15Further-\nmore, we point out that the temperature dependence\nof 1/T1Tis close to a T−4/3dependence below 10 K,\nwhich is an expected behavior when a 3-D metal is close\nto the FM instability.16The characteristic energy of\nthe spin fluctuation T0is estimated to be T0∼220 K\nfrom the Sommerfeld coefficient γ,5,17and according to\nthe SCR calculation, 1 /T1Tis close to a T−4/3depen-\ndence below T/T0<0.1.16Therefore, the observation of\n1/T1T∼T−4/3below 10 K is reasonable and suggests\nthatUCoGepossessesanFMquantumcriticalcharacter.\nAlthough a clear anomaly was observed in 1 /T1Tat\nTCurie, the NQR spectra of ν1andν2are almost un-\nchanged even at 1.5 K, as observed in Fig. 4, indicative\nof the absence of the static internal field down to 1.5\nK. This might be related to the FM quantum critical\ncharacter because it is anticipated that ferromagnetism\ndevelops gradually near the quantum critical point. By\nfurther lowering the temperature, new NQR signals af-\nfected by the static fields start to develop below 1 K.4 J. Phys. Soc. Jpn. Full Paper Author Name\nThe Co-NQR spectra at 95 mK shown in Fig. 4 (c) can\nbe understood with the superposition of two NQR spec-\ntra; one is the Co-NQR spectra without the internal field\n(“nonmagnetic” NQR signals), and the other is the Co-\nNQR spectra affected by the internal field at the Co site\n(“magnetic”NQR signal). The magnitude of the internal\nfields of the magnetic NQR signal is approximately 400\nOe, and its direction is almost perpendicular to the prin-\ncipal axis of the EFG. In Fig. 4, the calculated resonance\nfrequencies are shown by arrow.18The NQR spectra in\nthe FM state indicate that there exist two Co sites with\nand without the internal magnetic fields. Furthermore,\nthe temperature variation of the NQR spectrum in the\nFM state shows a first-order-like behavior because the\ninternal field of the magnetic NQR signal is independent\nof temperature and only a fraction of the magnetic and\nnonmagnetic NQR signals changes by due to tempera-\nture.\nNext, we discuss the SC properties revealed by the\n1/T1measurement. Below 0.7 K, which is the onset tem-\nperature of the SC transition, the longer component of\nT1ascribed to the opening of the SC gap was observed in\nm(t), as shown in the inset (c) of Fig. 5. The longer and\nshorter components of 1 /T1are shown in the main panel\nof Fig. 5. The longer component of 1 /T1is estimated\nto be 30% of the total relaxation component, which is\ndetermined by the extrapolation of m(t) tot= 0, and\nit follows a T3dependence below 0.3 K down to 0.1 K.\nTheT3dependence of 1 /T1suggests the presence of line\nnodes in the SC gap, and the complete Tdependence of\n1/T1belowTScan be interpreted by the 3-D polar state\nwith ∆(θ) = ∆ 0cosθ. TheTdependence of 1 /T1calcu-\nlated using the model is shown in Fig. 5 with the dotted\ncurve below TS, where the magnitude of the SC gap is\ntaken as 2∆ 0/kBTS= 4. The magnitude of the SC gap is\nquite comparable to that in UGe 2.6,8It should be noted\nthat an appreciable Korringa ( T1T= constant) behav-\nior was not observed down to 95 mK, indicative of the\nsmall effect of the residual density of state. In contrast,\nthe shorter component of 1 /T1shows a√\nTdependence\nbelowTS, suggestive of the persistence of the magnetic\nfluctuationsdownto95mK.Itisquiteunusualthatmag-\nnetic fluctuations continue to develop in the SC state.\nThe underlying problem is the relation between\nthe magnetism and superconductivity in UCoGe, i.e,\nwhether superconductivity occurs only in the region\nwithout the internal field. To clarify this, we measure\n1/T1at 95mKat variouslow-frequencyNQRpeaks.Fig-\nure 7 shows the recovery curves m(t) measured at differ-\nentfrequencies.Here,the peaksat4.83and5.25MHzare\nthe Co-NQR signals mainly arising from the region with-\nout the internal field, and the peaks at 4.49 and 4.9 MHz\nare the NQR signals appearing below TCurie, which are\naffected by the internal field, i.e. Co-NQR signals arising\nfrom the magnetic region. As observed in Fig. 7, m(t)\nmeasured at the 4 different peaks shows the same behav-\nior,indicatingthatthesuperconductingfractions(30%of\nthe total relaxation component) of the regions with and\nwithout the internal fields are not different. The same\nbehavior of m(t) at the 4 peaks excludes the possibility\nthat superconductivity occurs in one of the regions with0 50 100 150 200 0.01 0.1 1\n4.0 4.5 5.0 5.5 6.0 \n T=95mK \n 4.49 MHz \n 4.83 MHz \n 4.9 MHz \n 5.25 MHz m(t) = \nÝM\nÞ∞\nß \nà M(t) \ná/M \nâ∞\nã\n time (msec) T = 95 mK Intensity \näarb. units \nå\nFrequency (MHz) \nFig. 7. (Coloronline)Co nuclearrecoveries m(t) measured atvar-\niousfrequencies at95 mK.The insetshowsthe NQR-signalpea ks\nat 95 mK, where the recoveries are measured.\nand without the internal field; rather, superconductivity\noccurs in both regions from the NQR spectra. This sit-\nuation is different from the high-pressure ferromagnetic\nphase in UGe 2, where superconductivity is observed only\nin the ferromagnetic region.7We also point out that the\nshort 1/T1component is quite similar in both regions\nwith and without the internal field. This is considered to\nbe because the ordered moment is as small as 0.03 µBat\nT= 0 K.\nNext, we attempt to understand the presence of the\ntwo components of 1 /T1in the entire sample. We sug-\ngest the possibility ofthe self-induced vortex(SIV) state,\nwhich is considered to be realized when superconductiv-\nity occurs in the ferromagnetic state.19The condition for\nthe realization of the SIV state is Hc1<4πM < H c2,\nwhereMis the net magnetization of a ferromagnet.\nWhen this condition is satisfied, the vortices are gen-\nerated spontaneously. In this case, it is considered that\n1/T1intheSCstatehastwocomponentsarisingfromthe\nSIV and SC regions, respectively. Recently, we observed\ntwowell-separatedcomponentsof1 /T1intheregionclose\nto the SC vortices in LaRu 4P12.20\nAnother interesting possibility to be expected in\nUCoGe is the nonunitary spin-triplet state similar to\nthe A1 state in3He superfluidity, where the SC gap\nopens only in the up-spin band and not in the down-spin\nband.21In this case, it is considered that the Korringa\nbehavior with 1/4 of the magnitude of 1 /T1Tin the nor-\nmal state is anticipated to remain at low temperatures,\nwhich arises from the relaxation process occurring in the\ndown-spin intraband. This possibility has been discussed\nin UGe 2from the73Ge-NQR studies.8However, the ab-\nsence of such Korringa behavior in the low-temperature\nregion seems to exclude this possibility.\nIn conclusion, ferromagnetic and SC transitions were\nidentified at TCurie∼2.5 K and TS∼0.7 K in our\npolycrystalline sample. Ferromagnetic fluctuations thatJ. Phys. Soc. Jpn. Full Paper Author Name 5\npossess a quantum critical character are dominant above\nTCuriefrom the 1 /T1and Knight-shiftmeasurements. We\nfoundthat1 /T1equalto30%ofthe totalrelaxationcom-\nponent startstodecreasebelow TS, accompaniedbya T3\ndependence below 0.3 K, and that 1 /T1of the remaining\ncomponent shows a√\nTdependence, indicative of mag-\nnetic fluctuations even below TS. From the present NQR\nmeasurements in the SC state, we suggest that the self-\ninduced vortex state is realized in UCoGe.\nWe thank Y. Maeno and K. Yoshimura for experi-\nmental support and valuable discussions. We also thank\nN.Tateiwa,H.Kotegawa,Y.Itoh, C.Michioka,H. Ikeda,\nand S. Fujimoto for valuable discussions. This work was\npartiallysupportedbythe21stCOEprogramon“Center\nfor Diversity and Universality in Physics” by MEXT of\nJapan,and“Skutterudite”and “SuperClean”byMEXT\n(Nos.1827006,17071007)andGrants-in-AidforScientific\nResearch from the Japan Society for the Promotion of\nScience (JSPS)(No.18340102).\n1) S. S. Saxena et al.: Nature (London) 406(2000) 587.\n2) T. Akazawa et al.: J. Phys. Condens. Matter 16B (2004) L29.\n3) A. Huxley et al.: Phys. Rev. B 63(2001) 144519.4) D. Aoki et al.: Nature 413(2001) 613.\n5) N. T. Huy et al.: Phys. Rev. Lett. 99(2007) 067006.\n6) H. Kotegawa et al.: J. Phys. Soc. Jpn. 74(2005) 705.\n7) A. Harada et al.: J. Phys. Soc. Jpn. 74(2005) 2675.\n8) A. Harada et al.: Phys. Rev. B 75(2007) 140502.\n9) N. K. Sato: in preparation for publication.\n10) T. Ohta: in preparation for publication.\n11) M. Kontani, T. Hioki, and Y. Masuda: Solid.State. Commun .\n18(1976) 1251.\n12) M. Takigawa and H. Yasuoka: J. Phys. Soc. Jpn. 51(1982)\n787.\n13) K. Yoshimura et al.: J. Phys. Soc. Jpn. 53(1984) 503.\n14) T. Moriya and A. Kawabatra: J. Phys. Soc. Jpn. 34(1973)\n639.\n15) T. Moriya and A. Kawabatra: J. Phys. Soc. Jpn. 35(1973)\n669.\n16) A. Ishigaki and T. Moriya: J. Phys. Soc. Jpn. 65(1996) 3402.\n17) T. Moriya and T. Takimoto: J.Phys. Soc. Jpn. 64(1995) 960.\n18) For the calculation of the NQR frequencies affected by the\ninternal magnetic field, the effect of the internal field is\ntreated as perturbation. When the field was assumed to be\nH= (H0sinθcosφ,H0sinθsinφ,H0cosθ), the frequencies\nshown by the arrows are obtained by parameters of ( H0,θ,φ)\n= (400 Oe, 85◦, 45◦).\n19) M. Tachiki et al.: Solid State Commun. 34(1980) 19.\n20) Y. Nakai et al.: submitted to Phys. Rev. Lett.\n21) T. Ohmi and K. Machida: Phys. Rev. Lett. 71(1993) 625." }, { "title": "1601.03133v1.Spin_Vortex_Resonance_in_Non_planar_Ferromagnetic_Dots.pdf", "content": " \n \n Spin Vortex Resonance in Non -planar Ferromagnetic Dots \nJunjia Ding, Pavel Lapa,† Shikha Jain,‡ Trupti Khaire, Sergi Lendinez, Wei Zhang, Matthias B. \nJungfleisch, Christian M. Posada, Volodymyr G. Yefremenko,§ John E. Pearson, \nAxel Hoffmann, and Valentine Novosad* \nMaterials Science Division, Argonne National Laboratory, Argonne, IL 60439, USA \n \n \nAbstract \nIn planar structures, the vortex resonance frequency changes little as a function of an in -plane \nmagnetic field as long as the vortex state persists. Alte ring the topography of the element leads to \na vastly different dynamic response that arises due to the local vortex core confinement effect. In \nthis work, we studied the magnetic excitations in non -planar ferromagnetic dots using a \nbroadband microwave spec troscopy technique . Two distinct resonance frequency ranges were \nobserved depending on th e position of the vortex core controllable by applying a r elatively small \nmagnetic field. The micromagnetic simulations are i n qualitative agreement with the. \nexperime ntal results . \n \n† On leave from Texas A&M University, College Station, TX 77843, USA. \n‡ Presently at HGST, A Western Digital Company \n§ Presently at High Energy Physics Division, Argonne National Laboratory. \n* Correspondence and requests for materials should be addressed to V. N. (email: novosad@anl.gov) \n \n Introduction \nThe investigation of spin dynamics in geometrically confined ferromagnets is an important \nresearch topic due to both the technological and the basic science relevance. Some prominent \nexamples include the nanoscale spin textures with non -collinear magnetization arrangement \nsuch as skyrmions ,1 domain walls ,2,3 and spin vortices.4 A vortex -type magnetization is \nenergetically favorable for micron and sub -micron disks with no crystalline anisotropy. The \nlong-range dipolar forces govern the spins dy namics in the vortex -state. Of particular interest is a \nlow-frequency excitation mode associated with the vortex core gyration . Besides the \nmagnetization of saturation, the vortex eigen frequencies depend on the dot thickness -to-diameter \nratio.5 The excitation spectrum depends only weakly on the value of an in -plane applied field \nbecause the external field is compensated by the demagnetizing field of the shifted vortex.6 \nFurthermore, it has been reported that the intri nsic pinning due to grain boundaries,7 surface \nroughness8 or exchange bias9 could also influence the vortex core gyration frequency. While the \nvast majority of the works on the dynamics o f spin vortices deal with the geometrically flat \nelements, here we have studied the circular elements with intentionally altered topography. This \nwas achieved by using a pre -patterned substrate with the characteristic lateral dimensions \nsignificantly small er that the diameter of the desired magnetic element. In our case, the \nnonmagnetic disks were first defined via lift -off process, and then covered by another patterned \nferromagnetic layer with the same geometric center but a larger diameter. As the result, we \nobtained a “hat -like” structure of the sample, schematically depicted in Figure 1(a). Then the \ndynamic properties of these non -planar elements were systematically investigated using a \n \n broadband microwave spectroscopy and micromagnetic modeling as a fun ction of their \ndimensions and amplitude of the external field. As we show below, the introduced disk -shaped \nstep (also referred as a vortex barrier) provides a strong geometric confinement and vortex core \npinning effect. \n \nResults \nElement geometry. Two se ts of samples were fabricated. The first one is a reference sample, a \nflat Permalloy (Py, Ni 80Fe20 alloy) disk with diameter of 1 micron, and thickness of 50 nm. The \nsecond one is a non -planar, or an “engineered” dot, which is also circular in shape, with the same \ndimensions but has altered topography. Shown in Fig. 1(a) is a sketch of the cross -section profile \nof the engineered dot in which the magnetic material at the center has been lifted 25 nm by \npreparing a non -magnetic (Titanium) island prior to the deposition of the ferromagnetic material. \nFigures 1(b) and (c) show representative scanning electron microscope images of the non -planar \ndot with 200 nm diameter of the step and the reference dot. The step height Tb and the dot \nthickness TPy have been fixe d at 25 nm and 50 nm, respectively in all samples presented in this \npaper, while the inner diameter Db varied from 150 to 300 nm. \nDynamic response. In order to characterize the dynamic response of the samples, a microwave \ntransmission measurement has been performed in a broad frequency range using a vector \nnetwork analyzer (VNA). Applying an external magnetic field Happ allowed us to probe how the \n \n dynamic response of the vortex core changes once it has been displaced from the central area of \nthe disk. \nFigure 1(d) compares the absorption spectra of the engineered dot taken at remanence ( Happ \n= 0), and in applied field ( Happ = -180 Oe) after saturating the sample with a –1.5 kOe external \nfield. Interestingly, for a higher value of applied field we detect the resonance frequency of the \nvortex translation mode at ~0.4 GHz, while for the Happ = 0 Oe the frequency increases almost \n50%, to ~0.6 GHz. The same microwave measurements performed for a flat Ni 80Fe20 dot (e.g. \nthe reference sample, without vortex barrier ) are shown in Fig. 1(e). The frequency change when \nthe Happ is decreased from -180 Oe to 0 Oe is insignificant for the reference dot. A slightly higher \nfrequency for shifted vortex state is in fact expected. It originates from additional dipolar and \nexcha nge forces due the vortex spin structure deformation. The striking result is that while the \nresonance frequencies for the shifted vortex states are the same for both samples, their values in \nremanence differ significantly. This suggests that the dynamic re sponse of the engineered dot can \nbe fine -tuned by controlling the relative position of the vortex core with respect to the vortex \nbarrier using a relatively small magnetic field. One could speculate that for the engineered dot \nsubjected to Happ = -180 Oe, the vortex core is located outside the circumference of the barrier \nand thus it encounters the dipolar fields averaged within the outer boundary of the entire dot. As \nthe |Happ| value is reduced to zero, the vortex core shifts to the dot center and its gy ration is now \ngoverned by the dominant contribution from the magnetic charges confined within the barrier \nboundary. As a result, there appears the ~0.2 GHz frequency difference between these two core \n \n positions. Obviously, this effect is not observed for th e reference dot (Fig. 1(d)) due to the lack of \nthe barrier. \nTo further clarify the observed effect we fabricated and systematically investigated the dots \nwith variable barrier diameter. Figures 2 (b - e) are the representative absorption spectra taken at \nremanence ( Happ = 0 Oe) for the 50 nm thick, 1 micron diameter Py dots but with Db varied from \n150 to 300 nm. The vortex mode of all modified dots is characterized by the domination of one \npeak. The frequency of this main mode ( V0) decreases as the barrier diameter is increased. This \ntrend is similar to the results of previous studies10 for planar Ni 80Fe20 disks where the resonance \nfrequency was reported to decrease with the increase in the dot diameter (for a fixed thickness of \nthe element). The micromagnetic modeling yields similar results as it is discussed below. The 2D \nabsorption spectra (Fig. 2(a’ – e’)) also contain important information about the magnetization \nevolution process. Let us consider the reference dot ( Db = 0 nm, Fig. 2(a’)) and the dot with Db = \n200 nm (Fig. 2(c’)) as two examples. The vortex resonance corresponding to the translational \nmode appears at Happ ~250 Oe in both structures, indicating that the value of the vortex \nnucleation field is not affected by presence of the vortex barrier. Unlike to the almost field -\nindependent spectra of the reference dot, a discontinuous step -like frequency change is observed \nin the engineered dot. For instance, for Happ = -180 Oe and 0 Oe, the frequency differen ce is \n~200 MHz for the dot with Db = 200 nm (Fig. 1(d)), while it is only ~15 MHz for the reference \ndot (Fig. 1(e)). As it was mentioned before, this jump in the frequency is attributed to the \nchanges of the position of the vortex core. In a high magnetic field, when Happ is just below the \nnucleation field, the vortex core is located close to the outer edge. As the magnetic field Happ is \n \n gradually decreased, the vortex core progressively displaces towards to the center of the dot. The \nintensity of the reson ance line in an engineered dot remains unchanged till Happ = -150 Oe. In the \nfield range from -150 Oe to -60 Oe the signal disappears (or is below our experimental \nsensitivity limit). We speculate that in this fields range the vortex gyration is significan tly \nsuppressed due to the pinning effect of the barrier edges. (This vortex core pinning field range \nvaries as a function of the barrier diameter as the vertical line indicated in Fig. 2(b’ to e’).) The \nresonance reappears again when Happ is decreased belo w -60 Oe suggesting that the vortex core \nis now fully inside the barrier circumference. Stronger dynamic dipolar fields of the barrier cause \nthe gyrotropic mode frequency shift to a much higher values. The frequency does not change in \nsmall positive fields till the gyration stops when the core reaches the barrier edge again. With \nfurther magnetic field increase, the vortex core overcomes the barrier border and the low \nfrequency resonance line re -emerges again. Similarly, to the nucleation fields, the vortex \nannihilations fields almost do not differ in the modified and reference dots. Thus, while we have \nseen that altering the dot topography has a profound impact on the low -field vortex core \ndynamics, its overall magnetostatic properties (e.g. the hysteresis loop) remain unaffected. This \nis different from the case of dot -on-dot structure where bi -stable magnetic states were reported .11 \n \nMicromagnetic modeling: A systematic micromagnetic study has been performed to further \ninvestigate the static and dynamic response of the modified dots. The simulations confirm that in \nspite of such significant alteration to the topography o f the disk, the vortex state remains the \n \n ground state for the system. Figures 3 (a, b) show the magnetization distribution in remanence \nfor the dot with Db = 200 nm and in an in -plane magnetic field Happ = -180 Oe. \nA significant difference in the eigenfre quencies for these two distinct cases (e.g. the core \nlocated outside and inside of the barrier circumference) were also confirmed micromagnetically.. \nShown in Fig. 3(c) are the calculated relative energy profiles of the engineered dot (triangle \nsymbols) an d reference dot (square symbols) plotted as a function of the displacement of the \nvortex core. A clear difference between the two sets of results indicates the strong effect of the \nvortex confinement when the core is at the center of the element. The energ y profile vs the core \ndisplacement can be approximated as a parabolic function12 E(X) = E(0) +1/2X2, where is the \neffective stiffness coefficient, E(0) is the energy at the equilibrium position and X is the vortex \ncore displacement. It should be emphasized that the dot self -induced magnetostatic energy of \nmoving vortex provides the dominant contrib ution to E(X). Using the simulation data shown in \nFig. 3(c), one can find the remanent values of stiffness coefficients ( ) as 0.9410-20 J/nm2 and \n1.710-20 J/nm2 for the reference and engineered dots, respectively. Since the frequency is \ndirectly proport ional to ,13 the vortex core gyrates faster when trapped inside the barrier. There \nappears no significant difference between the energy profiles for engineered (tri angle symbols) \nand reference (square symbols) dots when the vortex core is located outside the barrier ( Happ = -\n180 Oe), Figure 3(d). The asymmetry in the energy profile due the vortex structure deformation \nin the shifted state can be accounted by adding a cubic term.6 \nAs the vortex core is very small, the dipolar forces originating from the dynamic magnetic \ncharges outside the vortex core govern its gyration. These cha rges can be calculated using a so - \n \n called “side -surface charges free” analytical model.14 Within this model, the magnetization \ndistribution of precessing vortex obeys boundary conditions such that there is no net \nmagnetization component perpendicular to the dot's lateral surface. Altering the topograph y that \nleads to formation of a step -like barrier will inevitably impose an additional requirement so the \nmagnetic “charges” on the barrier edges surfaces are minimized as well. Figure 3 (e) shows \nrepresentative images of temporal changes in the divergence of the magnetization for flat and \nengineered dots (upper and lower images, respectively). It is clear that the volume magnetostatic \ncharges contributing to the vortex dynamics and defining its eigenfrequency are well -distributed \nacross the dot for the ref erence sample, while their counterparts in the engineered dot are \npredominantly located in the central area circumscribed by the barrier edge. The same simulation \nwas also performed for Happ = -180 Oe as shown in Fig. 3(f). To our surprise, in this case th e \nspatial distribution of changes in the volume charges is almost identical for both samples \nresulting in a similar vortex gyration frequencies. \nFinally, to further understand the effect of the vortex barrier to the translational mode \nfrequency, systemat ic simulations have been performed as a function of the barrier size. Shown \nin Fig. 4(a) is the summary of the experimental (square symbols) and micromagnetic (solid line) \nfrequencies as a function of Db. While there is a noticeable quantitative discrepanc y between the \nexperimental and computational results, they are in a good qualitative agreement. Figure 4(b) \nshows the simulated frequency plotted as a function of barrier thickness Tb. The frequency \ncontinuously increases with Tb, it almost doubles in comp arison to the reference dot for Tb = 40 \nnm. These results demonstrate that the vortex resonance frequency can be effectively controlled \n \n by adjusting the barrier geometry. Interestingly, the low field experimental and micromagnetic \nfrequencies for engineere d dots shown in Fig. 3(a, b) scale universally when replotted as a \nfunction the barrier geometric aspect ratio Tb/Db. This is similar to how the translational mode \nfrequency in flat disks scales as a function of the disk thickness to diameter ratio.10 \n \nSummary \nA nonmagnetic nanodot inserted under a mesoscale Ni 80Fe20 dot was shown to provide a \ngeometric confinement effect causing the changes in the vortex translational mode frequency. \nTwo distinct resonance frequen cy ranges were observed depending on the position of the vortex \ncore (inside or outside of the barrier) controllable by applying a relatively small magnetic field. \nBy comparing the experimental data and micromagnetic simulations if was found that the \nfrequ ency of the gyrotropic mode increases as the thickness -diameter ratio of the barrier is \nincreased. Further studies of such non -planar ferromagnetic elements will be focused on the \ndetails of the pinning mechanism, its possible impact on the energetics of t he vortex core reversal \nprocess and the high frequency spin dynamics. \n \nMethods \nSample fabrication. The engineered dots were fabricated using a multistep electron -beam (EBL) \nlithography process. First, the disk arrays with diameter in a range of 150 nm to 3 00 nm and \nalignment marks were defined on polymethyl methacrylate (PMMA) resist, accompanied by e - \n \n beam evaporation and lift -off process of a 25 -nm-thick titanium film. The second step EBL \npatterning of 1 -micron diameter disks followed by deposition of 50 -nm-thick Ni 80Fe20 and lift -\noff completes the fabrication process. The barriers and the disk are concentric as is confirmed by \nScanning Electron Microscopy imaging. \n \nSpectral measurements. In order to characterize the dynamic properties of the samples, a \ncoplanar waveguide (CPW) with a 3 μm -wide -signal line was fabricated on top of each dot array \nusing optical lithography followed by a Ti(5 nm)/Au(150 nm) sputter deposition and a lift -off \nprocess. Microwave transmission measurements have been performed in th e 0.05 ~ 10 GHz \nfrequency range using a broadband microwave vector network analyzer (VNA). The microwave \ntransmission was measured by sweeping the frequency for fixed magnetic field. Applying an in \nplane magnetic field allowed us to probe how the dynamic r esponse of the vortex core changes \nonce it has been displaced from the central area of the disks. Since the focus of this paper is the \nfundamental gyrotropic vortex mode, all results are presented in the field range of -500 ~ 500 Oe \nand frequency range of 0.05 ~ 1.0 GHz . Prior to the magnetic field sweep, the samples were \nmagnetized at the 1.5 kOe field. \n \nMicromagnetic simulation details. Systematic micromagnetic modeling was performed using \nmumax3 code .15 Typical parameters for Ni 80Fe20 (the saturation magnetization Ms = 700 kA/m, \nthe exchange constant A = 13×10-12 J/m, the damping factor 0.01, the gyromagnetic ratio γ = 2.8 \n \n GHz/kOe, negligible magnetocrystalline anisotropy) and 5 ×5×5 nm3 cell size were used in the \nsimulation. \n \n References: \n1. Jiang, W. et al. Blowing magnetic skyrmion bubbles. Science (80 -. ). 349, 283–286 (2015). \n2. Parkin, S. S. P., Hayashi, M. & Thomas, L. Magnetic Domain -Wall Racetrack Memory. \nScience (80 -. ). 320, 190–194 (2008). \n3. Parkin, S. & Yang, S. -H. Memory on the racetrack. Nat. Nanotechnol. 10, 195–198 (2015). \n4. Guslienko, K. Y., Kakazei, G. N., Ding, J., Liu, X. M. & Adeyeye, A. O. Giant moving \nvortex mass in thick magnetic nanodots. Nat. Publ. Gr. 1–8 (2015). \ndoi:10.1038/srep13881 \n5. Novosad, V. et al. Magnetic vortex resonance in patterned ferromagnetic dots. Phys. Rev. \nB 72, 24455 (2005). \n6. Buchanan, K. S. et al. Magnetic -field tunability of the vortex translational mode in \nmicron -sized permalloy ellipses: Experiment and micromagnetic modeling. Phys. Rev. B \n74, 064404 (2006). \n7. Compton, R. L. & Crowell, P. a. Dynamics of a pinned magnetic vortex. Phys. Rev. Lett. \n97, 137202 (2006). \n8. Compton, R. L., Chen, T. Y. & Crowell, P. a. Magnetic vortex dynamics in the presence \nof pinning. Phys. Rev. B 81, 144412 (2010). \n9. Sort, J. et al. Imprinting Vortices into Antiferromagnets. Phys. Rev. Lett. 97, 067201 \n(2006). \n10. Guslienko, K., Novosad, V., Otani, Y., Shima, H. & Fukamichi, K. Magnetization reversal \ndue to vortex nucleation, displacement, and annihilation in submicron ferromagnetic dot \narrays. Phys. Rev. B 65, 024414 (2001). \n11. Stebliy, M. E. et al. High -frequency switching of magnetic bistability in an asymmetric \ndouble disk nanostructure. Appl. Phys. Lett. 104, 112405 (2014). \n12. Guslienko, K. Y. et al. Eigenfrequencies of vortex state excitations in magnetic \nsubmicron -size disks. J. Appl. Phys. 91, 8037 (2002). \n13. Guslienko, K. Y., Novosad, V., Otani, Y., Shima, H. & Fukamichi, K. Field evolution of \nmagnetic vortex state in ferromagnetic disks. Appl. Phys. Lett. 78, 3848 –3850 (2001). \n14. Metlov, K. L. & Guslienko, K. Y. Stability of magne tic vortex in soft magnetic nano -sized \ncircular cylinder. J. Magn. Magn. Mater. 242-245, 1015 –1017 (2001). \n15. Vansteenkiste, A. et al. The design and verification of MuMax3. AIP Adv. 4, 107133 \n(2014). \n \n \n Acknowledgements \nThis work was supported by the U. S . Department of Energy (DOE), Office of Science, Basic \nEnergy Sciences (BES), under Award # DE -AC02 -06CH11357. \n \n Figure Captions \nFig. 1 : (a) A sketch of the dot with Db = 200 nm with the simulated remnant magnetization state \nof the engineered dot. The x -component of the magnetization is represented from red to blue in a \nrange of +1 to -1. Scanning electron micrographs of the dot with (b) 200 nm barrier diameter ( Db) \nand (c) reference dot. (d) and (e) shows the experimental FMR curves with Happ = -180 Oe and 0 \nOe for Db = 200 nm and 0 nm, respectively \n \nFig. 2 : (a-e) The experimental remanent FMR absorption curves of the dots with Db = 0 nm \n(reference dot), 150 nm, 200 nm, 250 nm and 300 nm. The corresponding 2D FMR absorptions \nspectrums are shown in (a ’ to e’ ). The vertical lines in (b’ to e’) indicate the field range that \nvortex core jump is pinned at the barrier. \n \nFig. 3 : The simulated magnetization state (central layer) of the dot with Db = 200 nm for Happ = 0 \nOe (a) and -180 Oe (b). (c) and (d) shows the plots of the simulated energy profile of both the \nreference dot and the engineered dot as a function of the core displacement for Happ = 0 Oe and -\n180 Oe. Solid lines are the fitting results of the profile with the parabolic function. The simulated \nmagnetic charge density ( M) changes when the core is displaced from the equilibrium position \n(when it is performing gyrotropic movement) for the reference dot and the engineered dot with \nHapp = 0 Oe (e) and -180 Oe (f). The color code is in log scale. \n \n \nFig. 4 : (a) and (b) shows the simulated and experimental results of the gyrotropic mode \nfrequency of the modified dots as a function of the barrier diameter ( Db) and the barrier thickness \n(Tb), respectively. All results have been reformatted and plotted as a function of the barrier \naspect ratio ( Tb/Db) as shown in Fig. 4(c). \n \n \n \nFig. 1\n200 nmDb= 200 nmD= 1000 nm (b)\n(a)\nTb= 25 nm TPy= 50 nm\n \n0.2 0.4 0.6 0.8 1.0\n \nHapp= 0 OeHapp= -180 Oe(d)Absorpation (arb. unit)\n \n0.2 0.4 0.6 0.8 1.0\n \nHapp= 0 OeHapp= -180 Oe(e)\nFrequency ( GHz)Absorpation (arb. unit)Db= 0 nmDb= 200 nm\nV0V1\nV1\nV1\n200 nmD= 1000 nm (c) \n \n \n \nFig. 2Microwave Absorpation (arb. unit)\n (b)\n (a)\n \n(c)\n \n(d)\n0.2 0.4 0.6 0.8 1.0\n \n(e)Db= 150 nm\nDb= 200 nm\nDb= 250 nm\nDb= 300 nmV0V1\nHapp(Oe)\n0.20.40.60.81.0\n \n Db= 150 nm(b’)\n0.20.40.60.81.0\n \n \nDb= 200 nm(c’)\n0.20.40.60.81.0\n \n \nDb= 250 nm(d’)\n-500 -250 0 250 5000.20.40.60.81.0\n \n \nDb= 300 nm(e’)Frequency (GHz)\n0.20.40.60.81.0\n \n Db= 0 nm(a’) Db= 0 nm\nReference DotReference Dot\nFrequency ( GHz) \n \n \n \nFig. 3\n(a)Happ= 0 Oe\n(b)Happ= -180 Oe\n-15 -10 -5 0 5 10 15-50510152025\n \nCore Displacement ( X, nm)(d)ΔEnergy (10-17J)\n-15 -10 -5 0 5 10 15-50510152025\n (c)ΔEnergy (10-17J)\nCore Displacement ( X, nm)\n(f)\n Ref. Dot Db= 200 nm Ref. Dot Db= 200 nm\nhigh\nlow(e)\nRef. Dot\nDb= 200 nm\nRef. Dot\nDb= 200 nm \n \n \n \nFig. 4\n100 200 300 4000.30.40.50.60.70.8\n Vortex Barrier Diameter ( Db, nm)Frequency (GHz)Simulation Results\nExperimental Results(a)\nTb= 25 nm\nHapp= 0 Oe\n0 10 20 30 40\n \nVortex Barrier Thickness ( Tb, nm)(b) Simulation Results\nDb= 200 nm\nHapp= 0 Oe\nDbTb\n0.0 0.1 0.2 0.3\n \nBarrier Aspect Ratio ( Tb/Db)(c)\nFixed Db= 200 nm\nFixed Tb= 25 nm\nExperimental Results " }, { "title": "2303.05175v2.Antiferromagnetic_resonances_in_superconductor_ferromagnet_multilayers.pdf", "content": "arXiv:2303.05175v2 [cond-mat.supr-con] 24 Mar 2023Antiferromagnetic resonances in superconductor-ferroma gnet multilayers\nI. A. Golovchanskiy1,2,3,∗, V. V. Ryazanov1,2,4, V. S. Stolyarov1,2,3\n1Moscow Institute of Physics and Technology, State Universi ty,\n9 Institutskiy per., Dolgoprudny, Moscow Region, 141700,\nRussia;2National University of Science and Technology MISIS, 4 Leni nsky prosp.,\nMoscow, 119049, Russia;3Dukhov Research Institute of Automatics (VNIIA),\n127055 Moscow, Russia;4Institute of Solid State Physics (ISSP RAS),\nChernogolovka, 142432, Moscow region, Russia.\nIn this work, we study magnetization dynamics in supercondu ctor-ferromagnet (S-F) thin-film\nmultilayer. Theoretical considerations supported by the b road-band ferromagnetic resonance spec-\ntroscopy reveal development of acoustic and optic resonanc e modes in S-F multilayers at signifi-\ncantly higher frequencies in comparison to the Kittel mode o f individual F-layers. These modes are\nformed due to antiferromagnetic-like interaction between F-layers via shared circulating supercon-\nducting currents in S-layers . The gap between resonance mod es is determined by the thickness\nand superconducting penetration depth in S-layers. Overal l, rich spectrum of S-F multilayers and\nits tunability opens wide prospects for application of thes e multialyers in magnonics as well as in\nvarious superconducting hybrid systems.\nIntroduction. Hybridization of antagonistic super-\nconducting (S) and ferromagnetic (F) orders offers in\nelectronics and spintronics, which have been repeatedly\ndemonstratedinpastdecades[1]. Recentlytheinterestin\nS-F hybridization has been reinforced by demonstrations\nof its prospects in relation with the magnetization dy-\nnamics phenomena. In particular, interactions between\nmagnetization dynamics and the superconducting vortex\nlattice allow to form and guide the tunable magnonic\nband structure [2], as well as to induce exchange spin\nwaves by the DC electric current [3]. Also, interactions\nbetween magnetization dynamics and superconducting\nMeissner currents in hybrid structures modifies the spin-\nwave dispersion [4, 5], which can be used for creation\nof magnonic crystals [6] or for gating magnon currents\n[7]. Remarkably, low speed of electromagnetic propa-\ngation in superconductor-insulator-superconductor thin-\nfilm structures facilitates achievement of the ultra-strong\nphoton-to-magnon coupling in on-chip hybrid devices\n[8, 9] aiming forthe photon-to-magnonentanglement[10].\nA new strong phenomenon in S-F hybrid structures\nwas reported recently in Refs. [11–13] and investigated\nfurther in Refs. [14, 15]. In superconductor-ferromagnet-\nsuperconductor (S-F-S) thin-film structures in the pres-\nence of electronic interaction between superconducting\nand ferromagnetic layers a radical increase in the ferro-\nmagnetic resonance (FMR) frequency occurs. The mech-\nanism behind the phenomenon constitutes a formation\nof one-dimensional superconducting torque via the in-\nterplay between the superconducting imaginary conduc-\ntance and magnetization precession at S-F interfaces,\nwhich result in induction ofalternatingcirculatingsuper-\nconducting currents in the opposite phase to the magne-\ntization precession.\nIn this work, we generalize the problem and consider\nmagnetization dynamics in arbitrary S-F multilayers.\nCouplingbetweenferromagneticlayersviasuperconduct-\ning currents allows to induce antiferromagnetic-like in-\nteraction between F-layers, which result in acoustic andoptic resonances modes. Theoretical considerations sup-\nported by the broadband ferromagnetic resonance spec-\ntroscopy demonstrate that the spectrum is determined\nby geometrical characteristics of a multilayer as well as\nby the superconducting penetration depth in S-layers.\nTheory. Following Refs. [15, 16], electrodynamics and\nmagnetizationdynamicsinS-Fmultilayersobeysconven-\ntional Maxwell equations supplemented by the Ohm law\nwith imaginary conductance in superconducting layers\nand by the Polder susceptibility in ferromagnetic layers.\nBy neglecting edge effects, the y-component of magnetic\nfield as well as x-components of electric field and of the\ncurrent are functions of the transverse coordinate zonly\n(see Fig. 1). Derivation of the magnetic field in S- and\nF-layers from initial Maxwell equations yields following\ngeneral expressions\nHS\ny(z) =Aiexpz\nλS+Biexp−z\nλS,\nHF\ny(z) =Ciexpz\nλF+Diexp−z\nλF,(1)\nwhere the subscript of coefficients ispecifies the su-\nperconducting layer or the ferromagnetic layer in the\nstack,λSis the superconducting penetration depth, and\nλF=δFΩ is the ferromagnetic penetration depth, δF=/radicalbig\ni/µ0ωσFis the conventional electromagnetic penetra-\ntion depth into a metal with conductivity σF(typically,\nin permalloy σF∼106Ohm−1m−1). The characteristic\ndimensionless frequency Ω is given by\nΩ2=γ2(H+Ha)(H+Ha+Meff)−ω2\nγ2(H+Ha+Meff)2−ω2,(2)\nwhereHis the external field (aligned with the x-axis\nin Fig. 1),Hais the effective field of the uniaxial\nanisotropy aligned with the external field, and Meff=\nMs−2Ku/µ0Msis the effective magnetization, which ac-\ncounts the out-of-plane uniaxial anisotropy with the con-\nstantKu. Notice that the conventional Kittel formula2\nfor the ferromagnetic resonance in thin films in these no-\ntations is provided by Ω = 0. At every S/F interface the\nfollowing boundary conditions are fulfilled\nHS\ny=HF\ny,\n1\nσSdHS\ny\ndz=1\nσFdHF\ny\ndz,(3)\nwhereσS=i/µ0ωλ2\nSis the imaginary conductance in\nsuperconducting layers.\nFIG. 1. Schematic illustration of the interplay between ac\nmagnetic field, magnetization precession and superconduct -\ning currents in S-F-S trilayer. Magnetization precession ( /vectorM,\nblack arrow) at S-F interfaces induces macroscopic superco n-\nducting currents alternating in S-layers along the x-direction\n(JS\nx, blue arrows). These currents form the magnetic field HS\ny\nin the F-layer along the y-direction in opposite phase to the\nprecession of /vectorM.\nFor the S-F-S trilayer, depicted schematicallyin Fig. 1,\nsolution of Eq. 3together with natural boundary condi-\ntions at outer surfaces of S-layers, namely, HS\ny= 0, and\nin the limit dF≪λFyields the following expression for\nferromagnetic resonance frequency:\nΩ2=−dF\nλStanhdS1/λStanhdS2/λS\ntanhdS1/λS+tanhdS2/λS.(4)\nAs example, black and red curves in Fig. 2compares the\ndependence of the resonance frequency on the magnetic\nfieldfr(H), respectively, in a single F-layer and in S-F-\nS trilayer with the following thicknesses: dF1= 50 nm,\ndS1= 150 nm, dS2= 100 nm, λS= 100 nm.\nApplication of the same derivation approach for the\nsymmetric S-F-S-F-S multilayer, depicted schematically\nin Fig.3, yields two resonance modes with the in-phase\n(acoustic mode) and the anti-phase (optic mode) preces-\nsion of ferromagnetic layers, respectively,\nΩ2\na=−dF\nλStanhdSi/λScothdSe/2λS\ntanhdSi/λS+cothdSe/2λS,\nΩ2\no=−dF\nλStanhdSi/λStanhdSe/2λS\ntanhdSi/λS+tanhdSe/2λS,(5)\nwhereSedenotes external superconducting layers (S1\nand S3 in Fig. 3),Sicorresponds to the internal super-\nconducting layer (S2 in Fig. 3), and thicknesses of both\nferromagneticlayersis consideredequal dF1=dF2=dF.\nBlue curves in Fig. 2show acoustic and optic resonance\ncurvesfr(H) in the multilayer with the same thicknesses/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56/s82/s101/s115/s111/s110/s97/s110/s99/s101/s32/s102/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102\n/s114/s44/s32/s71/s72/s122\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84\nFIG. 2. Theoretical dependencies of the resonance frequenc y\non the magnetic field fr(H) in conventional F-layer (solid\nblack curve, Kittel formula), S-F-S trilayer (solid red cur ve,\nEq. 4) and symmetric S-F-S-F-S multilayer (solid and dashed\nblue curves, Eq. 5). The following parameters are used for\ncalculations: γ/2π= 29.5 GHz/T, dF1=dF2= 50 nm,\ndS1=dS3=dSe= 150 nm, dS2=dSi= 100 nm,\nλS= 100 nm, µ0Meff= 1 T,Ha= 0.\nas in a single F-layer and in S-F-S trilayer : dF1=dF2=\ndF= 50 nm, dS1=dS3=dSe= 150 nm, dSi= 100 nm,\nλS= 100 nm.\nFIG. 3. Schematic illustration of the interplay between ac\nmagnetic field, magnetization precession andsuperconduct ing\ncurrents in symmetric S-F-S-F-S multilayer. Acoustic (a) a nd\noptic (b) modes a formed.\nIt should be noticed that the acoustic mode (Fig. 3a)\ncan be thought as being formed by the global supercon-\nducting current, which circulates in external supercon-\nducting layers S1 and S3, while the internal S-layer S2\nonly screens the induced magnetic field in conventional\nmanner. In fact, in the limit dSe→ ∞the expression\nfor Ωain Eq.5meets Ω in Eq. 4with the substitution\ndF→2dF. The optic mode (Fig. 3b) can be thought as\nthe resonance in S-F-S trilayer with reduced thickness of\ninternal S-layer: the expression for Ω oin Eq.5meets Ω3\n/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102/s44/s32/s71/s72/s122\n/s45 /s48/s46/s48/s49/s48\n/s45 /s48/s46/s48/s48/s54\n/s45 /s48/s46/s48/s48/s50\n/s48/s46/s48/s48/s50\n/s48/s46/s48/s48/s54\n/s48/s46/s48/s49/s48\n/s40/s97/s41/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s40/s98/s41/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102/s44/s32/s71/s72/s122\n/s45 /s48/s46/s48/s49/s48\n/s45 /s48/s46/s48/s48/s54\n/s45 /s48/s46/s48/s48/s50\n/s48/s46/s48/s48/s50\n/s48/s46/s48/s48/s54\n/s48/s46/s48/s49/s48\n/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s40/s99/s41/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84/s70/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102/s44/s32/s71/s72/s122\n/s45 /s48/s46/s48/s53\n/s45 /s48/s46/s48/s51\n/s45 /s48/s46/s48/s49\n/s48/s46/s48/s49\n/s48/s46/s48/s51\n/s48/s46/s48/s53\n/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s40/s100/s41/s32/s79/s112/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s101/s120/s112\n/s32/s79/s112/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s115/s105/s109\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s101/s120/s112\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s115/s105/s109/s82/s101/s115/s111/s110/s97/s110/s99/s101/s32/s102/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102/s44/s32/s71/s72/s122\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84/s48/s46/s48/s48 /s48/s46/s48/s51 /s48/s46/s48/s54 /s48/s46/s48/s57 /s48/s46/s49/s50 /s48/s46/s49/s53 /s48/s46/s49/s56 /s48/s46/s50/s49/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52/s49/s54/s49/s56\n/s32/s79/s112/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s101/s120/s112\n/s32/s79/s112/s116/s105/s99/s32/s109/s111/s100/s101/s44/s32/s115/s105/s109\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s32/s49/s44/s32/s101/s120/s112\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s32/s49/s44/s32/s115/s105/s109\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s32/s50/s44/s32/s101/s120/s112\n/s32/s65/s99/s111/s117/s115/s116/s105/s99/s32/s109/s111/s100/s101/s32/s49/s44/s32/s115/s105/s109/s82/s101/s115/s111/s110/s97/s110/s99/s101/s32/s102/s114/s101/s113/s117/s101/s110/s99/s121/s32 /s102/s44/s32/s71/s72/s122\n/s77/s97/s103/s110/s101/s116/s105/s99/s32/s102/s105/s101/s108/s100/s32\n/s48/s72 /s44/s32/s84 /s40/s101/s41\nFIG. 4. a-c) Differentiated transmission spectra dS21/dH(f,H) for samples SF1 (a) and SF2 (b) at temperature 2 K and\nfor sample SF1 at 8 K (c). Colour codes are provided in insets. d,f) Experimental (symbols) and theoretical (solid curves )\nresonance lines for samples SF1 (a) and SF2 (b) at temperatur e 2 K.\nin Eq.4with the substitution dS2=dSi/2. Qualitatively\nit can be concluded that the optic mode is unaffected by\ntheinteractionbetweenF-layerslayers,whiletheacoustic\nmodegainsenergyduetothecoupling. Interestingly,this\nqualitative picture is in direct contradiction with magne-\ntizationdynamicsinexchange-coupledferromagneticlay-\ners [17–21], where regardless the details of the exchange\ninteraction the acoustic mode corresponds to magnetiza-\ntion dynamics in non-interacting magnetic layers. Also,\naccording to Fig. 2and Eq.5the optic mode is observed\nat lower frequencies in comparison to the acoustic mode,\nwhich characterises the coupling between ferromagnetic\nlayers via superconducting layers as antiferromagnetic.\nIn a way, such interaction between ferromagnetic lay-\ners via superconducting currents in adjacent layers re-\nminds interaction of fluxons in superconductor-insulator\nJosephson junction stacks [22–24].\nIn the general case, resonance modes of an arbitrary S-\nF multilayer, whichconsistof Nferromagneticlayersand\nN+1 superconducting layers, can be derived numerically\nfrom the set of equations 3in the matrix form, [ M]×\n[An,Bn,Cn,Dn]T= 0, by finding frequencies ωrthat\nobey the expression\ndet[M(ωr)] = 0. (6)\nExperimental details and results. Experimentallymag-\nnetization dynamicsin S-F multilayersisstudied bymea-\nsuring the ferromagnetic resonance absorption spectrum\nusing the VNA-FMR approach[25–27] and the samechip\nlayout and experimental setup as in Refs [13, 14]. A\nseries of niobium-permalloy(Py=Fe 20Ni80)-niobium (Nb-\nPy) multilayered structures are placed directly on topof the central transmission line of superconducting Nb\nwaveguide. Deposition of Nb-Py multilayers is per-\nformed in a single vacuum cycle ensuring the electron\ntransparency at Nb-Py interfaces. Thin Si or AlO x\nspacing layer is deposited between Nb co-planar waveg-\nuide and the multialyers in order to ensure electrical\ninsulation of the studied samples from the waveguide.\nTwo test samples have been studied: a sample with\ntwo ferromagnetic layers, that consist of Nb-101nm/Py-\n11nm/Nb-41nm/Py-11nm/Nb-41nm, referred to as SF2\nand a sample with three ferromagnetic layers, that\nconsist of Nb-101nm/Py-11nm/Nb-40nm/Py-11nm/Nb-\n40nm/Py-12nm/Nb-41nm, referred to as SF3 . The SF2\nsample is made asymmetric on purpose in order to pro-\nvide a finite dynamic susceptibility of the optic mode,\nwhich otherwise is zero and, thus, does not couple to uni-\nform microwave magnetic field of the transmission line.\nMicrowave spectroscopy of samples was performed\nby measuring the transmission characteristics S21(f,H)\nin the closed-cycle cryostat Oxford Instruments Triton\n(base temperature 1.2 K) equipped with the home-made\nsuperconducting solenoid. Spectroscopy was performed\nin the field range from -0.22 T to 0.22 T, in the fre-\nquency range from 0 up to 20 GHz, and in the tempera-\nture range from 2 to 11 K. Magnetic field was applied in-\nplane along the direction of the waveguide(see Ref. [14]).\nFMR spectra at different temperatures were analysed by\nfittingS21(f) characteristics at specified HandTwith\nthe Lorentz curve and, thus, obtaining the dependencies\nof the resonance frequency on magnetic field fr(H).\nFigure4a-c demonstrates the essence of the studied\nphenomenon: at temperatures below the critical tem-4\nperature of Nb, T < T c, the transmission spectrum for\nSF2 sample consist of two spectral lines (Fig. 4a) and\nfor SF3 sample consist of three spectral lines (Fig. 4b).\nAtT > T c(Fig.4c) FMR spectrum for both samples is\nreduced to a single spectral line, which obeys the con-\nventional Kittel formula (Eq. 2, Ω = 0). For both sam-\nples the fit of FMR curves at T > T cyields negligible\nanisotropy field µ0Ha≈2 mT, the effective magnetiza-\ntionµ0Meff≈1.108T, which is close to typical values of\nthe saturation magnetization of permalloy µ0Ms≈1 T,\nand no noticeable dependence of HaandMeffon tem-\nperature. Temperature dependencies of FMR spectra\nfor both samples yield superconducting critical temper-\naturesTc= 7.7 K for SF2 sample and Tc= 7.9 K for\nSF3 sample. The critical temperature of Nb layers is re-\nduced in comparison to the bulk critical temperature of\nNbTc≈9 K owing to the inverse proximity effect [28].\nAtT < T c(Fig.4a,b) FMR spectrum shifts to higher\nfrequencies and splits to spectral lines in accordance to\nthe number of F-layers in the stack. The strongest line,\nthe acoustic mode, is observed at the highest frequencies,\nwhile weaker lines at lower frequencies correspond to op-\ntic modes. Resonance lines were modelled with Eq. 6\nusingλSas the fitting parameter (see Fig. 4d,e). The\noptimum fit is obtained with λS= 115 nm for SF2 sam-\nple andλS= 98 nm for SF3 sample. The obtained λS\nis slightly higher than typical values in bulk Nb (about80 nm) due to the inverse proximity effect [15, 28]. A\nbetter fit could be obtained by considering a variation of\nλSin different S-layers. Thus, the provided theoretical\ndescription of the magnetization dynamics phenomenon\nin arbitrary S-F multilayers is verified.\nConclusion. Summarising,wereportastudyofmagne-\ntization dynamics in S-F multialyers. Theoreticalconsid-\nerations supported by experiments in a wide frequency,\nfield, and temperature ranges show that the coupling\nbetween ferromagnetic layers via superconducting layers\nresults in formation of antiferromagnetic interaction be-\ntween F-layers with the strength that depends of thick-\nness and superconducting properties of S-layers. This in-\nteraction between ferromagnetic layers is formed via su-\nperconducting currents and result in formation of acous-\ntic and optic spectral branches. These results open wide\nprospects for application of S-F multialyers in magnon-\nics and also bridges magnetization dynamics phenomena\nwith various superconducting circuits [29–31], hybrid de-\nvices [8, 9], and metamaterials [32]. Moreover, resonance\nproperties of S-F multilayers by changing the supercon-\nducting state of S-layers optically[15, 33, 34] or via elec-\ntric currents.\nAcknowledgements The authors acknowledge Dr M.\nSilaev for fruitful discussions. The research study was\nfinancially supported by the Russian Science Foundation\n(grant N 22-22-00314).\n[1]J. Linder and J. W. A. Robinson, Superconducting spin-\ntronics, Nat. Phys. 11, 307 (2015).\n[2]O. V. Dobrovolskiy, R. Sachser, T. Bracher, T. Fischer,\nV. V. Kruglyak, R. V. Vovk, V. A. Shklovskij, M. Huth,\nB. Hillebrands, and A. V. Chumak, Magnon-fluxon inter-\naction in a ferromagnet/superconductor heterostructure,\nNat. Phys. 15, 477 (2019).\n[3]O. V. Dobrovolskiy, Q. Wang, D. Y. Vodolazov,\nB. Budinska, R. Sachser, A. V. Chumak, M. Huth, and\nA. I. Buzdin, arXiv , 2103.10156 (2021).\n[4]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nV. V. Bolginov, V. V. Ryazanov, A. A. Golubov, and\nA. V. Ustinov, Adv. Funct. Mater. 28, 1802375 (2018).\n[5]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nV.V.Ryazanov, A.A.Golubov, andA.V.Ustinov,Mod-\nified dispersion law for spin waves coupled to a supercon-\nductor, J. Appl. Phys. 124, 233903 (2018).\n[6]I. A. Golovchanskiy, N. N. Abramov, V. S. Stol-\nyarov, P. S. Dzhumaev, O. V. Emelyanova, A. A.\nGolubov, V. V. Ryazanov, and A. V. Ustinov, Ferro-\nmagnet/superconductor hybridmagnonic metamaterials,\nAdv. Sci. 6, 1900435 (2019).\n[7]T. Yu and G. E. W. Bauer, Phys. Rev. Lett. 129, 117201\n(2022).\n[8]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nM. Weides, V. V. Ryazanov, A. A. Golubov, A. V.\nUstinov, and M. Y. Kupriyanov, Science Advances 7,\neabe8638 (2021).\n[9]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nA. A. Golubov, M. Y. Kupriyanov, V. V. Ryazanov,and A. V. Ustinov, Physical Review Applied 16, 034029\n(2021).\n[10]M. A. Silaev, Ultrastrong magnon-photon coupling and\nentanglement in superconductor/ferromagnet nanostruc-\ntures, arXiv:2211.00462 (2022).\n[11]L.-L. Li, Y.-L. Zhao, X.-X. Zhang, and Y. Sun, Possible\nevidence for spin-transfer torque induced by spin-triplet\nsupercurrents, Chin. Phys. Lett. 35, 077401 (2018).\n[12]K.-R. Jeon, C. Ciccarelli, H. Kurebayashi, L. F. Cohen,\nX. Montiel, M. Eschrig, T. Wagner, S. Komori, A. Sri-\nvastava, J. W. Robinson, and M. G. Blamire, Effect of\nmeissner screening and trapped magnetic flux on magne-\ntization dynamics in thicknb/ni 80fe20/nb trilayers, Phys.\nRev. Appl. 11, 014061 (2019).\n[13]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nV. I. Chichkov, M. Silaev, I. V. Shchetinin, A. A.\nGolubov, V. V. Ryazanov, A. V. Ustinov, and M. Y.\nKupriyanov, Phys. Rev. Appl. 14, 024086 (2020).\n[14]I. A. Golovchanskiy, N. N. Abramov, O. V. Emelyanova,\nI. V. Shchetinin, V. V. Ryazanov, A. A. Golubov, and\nV. S. Stolyarov, Physical Review Applied 19, 034025\n(2023).\n[15]M. A. Silaev, Physical Review Applied 18, L061004\n(2022).\n[16]M. Kostylev, Journal of Applied Physics 106, 043903\n(2009).\n[17]D.S.Schmool andJ. M.Barandiaran, J. Phys.: Condens.\nMatter10, 10679 (1998).\n[18]J. LindnerandK.Baberschke,J. Phys.: Condens.Matter\n15, R193 (2003).5\n[19]S. M. Rezende, A. Azevedo, and R. L. Rodriguez-Suarez,\nJournal of Applied Physics 126, 151101 (2019).\n[20]I. A. Golovchanskiy and V. S. Stolyarov, Journal of Ap-\nplied Physics 131, 053901 (2022).\n[21]I. A. Golovchanskiy et al., Phys. Rev. B 106, 024412\n(2022).\n[22]T. Holst, J. B. Hansen, N. Grnbech-Jensen, and J. A.\nBlackburn, Phys. Rev. B 42, 127 (1990).\n[23]S. Sakai, P. Bodin, and N. F. Pedersen, Journal of Ap-\nplied Physics 73, 2411 (1993).\n[24]A. V. Ustinov and H. Kohlstedt, Phys. Rev. B 54, 6111\n(1996).\n[25]I. Neudecker, G. Woltersdorf, B. Heinrich, T. Okuno,\nG. Gubbiotti, and C. Back, Comparison of frequency,\nfield, and time domain ferromagnetic resonance methods,\nJ. Magn. Magn. Mat. 307, 148 (2006).\n[26]S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L.\nSchneider, P. Kabos, T. J. Silva, and J. P. Nibarger,\nFerromagnetic resonance linewidth in metallic thin films:\nComparison of measurement methods, J. Appl. Phys. 99,\n093909 (2006).[27]Y.-C. Chen, D.-S. Hung, Y.-D. Yao, S.-F. Lee, H.-P. Ji,\nand C. Yu, Ferromagnetic resonance study of thickness-\ndependent magnetization precession in ni 80fe20films, J.\nAppl. Phys. 101, 09C104 (2007).\n[28]J. Aarts, J. M. E. Geers, E. Br ´’uck, A. A. Golubov, and\nR. Coehoorn, Phys. Rev. B 56, 2779 (1997).\n[29]S. E. Barnes, M. Aprili, I. Petkovic, and S. Maekawa,\nSupercond. Sci. Technol. 24, 024020 (2011).\n[30]S. Mai, E. Kandelaki, A. F. Volkov, and K. B. Efetov,\nPhys. Rev. B 84, 144519 (2011).\n[31]I. A. Golovchanskiy, N. N. Abramov, V. S. Stolyarov,\nO. V. Emelyanova, A. A. Golubov, A. V. Ustinov, and\nV. V. Ryazanov, Supercond. Sci. Technol. 30, 054005\n(2017).\n[32]A. Pimenov, A. Loidl, P. Przyslupski, and B. Dabrowski,\nPhys. Rev. Lett. 95, 247009 (2005).\n[33]I. S. Veshchunov, W. Magrini, S. V. Mironov, A. G.\nGodin, J.-B. Trebbia, A. I. Buzdin, P. Tamarat, and\nB. Lounis, Nat. Comm. 7, 12801 (2016).\n[34]W. Magrini, S. V. Mironov, A. Rochet, P. Tamarat, A. I.\nBuzdin, and B. Lounis, Appl. Phys. Lett. 114, 142601\n(2019)." }, { "title": "0906.4863v1.Conductance_oscillation_due_to_the_geometrical_resonance_in_FNS_double_junctions.pdf", "content": "arXiv:0906.4863v1 [cond-mat.mes-hall] 26 Jun 2009Conductance oscillation due to the geometrical resonance i nFNSdouble junctions\nHiroyuki Ohtori1,2, Hiroshi Imamura2\n1Institute of Applied Physics, University of Tsukuba, Tsuku ba 305-8573, Japan\n2NRI-AIST, Central 2, 1-1-1 Umezono, Tsukuba 305-8568, Japa n\nWe theoretically analyzed the Andreev reflection in ferroma gnetic metal / nonmagnetic metal\n/ superconductor double junctions with special attention t o the electron interference effect in the\nnonmagnetic metal layer. We showed that the conductance osc illates as a function of the bias\nvoltage due to the geometrical resonance. We found that the e xchange field and therefore the\nspin polarization of the ferromagnetic metal can be determi ned from the period of the conductance\noscillation, which is proportional to the square-root of th e exchange field.\nRecently much attention has been focused on the An-\ndreev reflection (AR) in ferromagnetic metal (FM) / su-\nperconductor (SC) contacts1,2,3,4,5,6,7,8,9,10since the spin\npolarization of conduction electrons is measured through\nthe suppression of the conductance below the supercon-\nducting gap. This method is called point contact An-\ndreev reflection (PCAR) spectroscopy.\nOn the other hand, the quasiparticle (QP) interfer-\nence in nonmagnetic metal(NM) / SC junctions has\nbeen extensively studied in the past11,12,13,14,15,16,17,18.\nAs shown in Refs.12,13,14, the interference of QPs in\nthe SC layer brings about the oscillation of the den-\nsity of states against the QP energy, which is called a\nTomasch oscillation. The interference in the NM layer\nalso brings about the oscillation of the density of states\nin the NM layer11,15, which is known as the deGennes-\nSaint-James bound state or the McMillan-Rowell oscilla-\ntion. NesherandKorenmeasuredthedynamic resistance\nof YBa 2Cu3O6.6/ YBa 2Cu2.55Fe0.45Oy/ YBa 2Cu3O6.6\njunctions and determined the renormalized Fermi veloc-\nity of QPs in the YBa 2Cu2.55Fe0.45Oylayer from the pe-\nriod of the McMillan-Rowell oscillation16.\nIn this paper, we theoretically analyze the Andreev re-\nflection in a FM/NM/SC double junction system with\nspecial attention to the electron interference effect in the\nNM layer. Following the work of Blonder, Tinkham, and\nKlapwijk (BTK)19, we solve the Bogoliubov-de Gennes\n(BdG)20equations and calculate the conductance. We\nshow that the conductance due to the Andreev reflection\noscillates as a function of the bias voltage because of the\ngeometrical resonance predicted by deGennes and Saint-\nJames. We obtain the analytical expression of the prob-\nability of the Andreev reflection under the Andreev ap-\nproximation and find that the period of the conductance\noscillation is proportional to the square-root of the ex-\nchange field. Therefore, we can determine the exchange\nfield and therefore the spin polarization of the FM layer\nfrom the period of the conductance oscillation.\nThe system we consider is comprised of the\nFM/NM/SC double junctions shown in Fig. 1(a). The\ncurrentflows alongthe x-axis, and the interfacesbetween\nFM/NM and NM/SC are located at x= 0 and x=d,\nrespectively. The system is described by the followingBdG equation20:\n/parenleftbigg\nH0−h(x)σ∆(x)\n∆∗(x)−H0−h(x)σ/parenrightbigg/parenleftbigg\nfσ(r)\ngσ(r)/parenrightbigg\n=E/parenleftbigg\nfσ(r)\ngσ(r)/parenrightbigg\n,\n(1)\nwhereH0≡ −(/planckover2pi12/2m)∇2+V(x)−µFis the single parti-\ncle Hamiltonian, Eis the QP energy measured from the\nFermi energy µF,V(x) is the interfacial barrier21, and\nσ= +(−) represents the up-(down-)spin band. The ex-\nchange field function h(x) is given by h(x) =h0[1−Θ(x)]\nwhereh0represents the exchange field in the FM layer\nand Θ(x) is the Heaviside step function. We employed\nthe two-band Stoner model for the FM layer for simplic-\nity. The superconducting gap function is expressed as\n∆(x) = ∆ 0Θ(x−d), where ∆ 0represents the supercon-\nducting gap in the SC layer. We assume that the system\nFIG. 1: (a) Schematic diagram of aFM/NM/SC double junc-\ntion. An NM with a thickness of dis sandwiched by FM and\nSC layers. (b) Schematic diagrams of energy vs. momen-\ntumof the FM/NM/SC double junction for aspin-upincident\nelectron are shown. The open circles denote holes, the filled\ncircles electrons, and the arrows point in the direction of t he\ngroup velocity. The incident electron with up-spin is denot ed\nby 0, along with the resulting scattering processes: Andree v\nreflection (1), normal reflection (2) at the FM/NM interface,\ntransmission to the NM (3, 4) and reflection at the NM/SC\ninterface (5, 6), transmission as a electron-like quasi-pa rticle\nto the SC (7) and that as a hole-like quasi-particle (8). (c)\nSchematic diagrams of energy vs. momentum in the FM layer\nfor a spin-down incident electron are shown.2\nhas translational symmetry in the transverse ( yandz)\ndirection, and therefore the wave vector parallel to the\ninterface k/bardbl≡(ky,kz) is a conserved quantity.\nThe general solutions of the BdG equation (1) in the\nFM (NM) layer are given by\nΨ±k+\nFM(NM) ,σ(r) =/parenleftbigg\n1\n0/parenrightbigg\ne±ik+\nFM(NM) ,σxSk/bardbl(r/bardbl),(2)\nΨ±k−\nFM(NM) ,σ(r) =/parenleftbigg\n0\n1/parenrightbigg\ne±ik−\nFM(NM) ,σxSk/bardbl(r/bardbl),(3)\nwhere S k/bardbl(r/bardbl) represents the eigen function in the trans-\nverse direction in the k/bardblchannel and k+(−)\nFM(NM) ,σis thex\ncomponent of the wave number of an electron (hole) with\nσ-spin defined as k±\nFM,σ=√\n2m\n/planckover2pi1/radicalbigµF±E+σh0−E/bardbl\nandk±\nNM=√\n2m\n/planckover2pi1/radicalbigµF±E−E/bardbl,whereE/bardbl=/planckover2pi12\n2mk2\n/bardbl. In\nthe SC layer, we have\nΨ±k+\nSC(r) =/parenleftbigg\nu0\nv0/parenrightbigg\ne±ik+\nSCxSk/bardbl(r/bardbl), (4)\nΨ±k−\nSC(r) =/parenleftbigg\nv0\nu0/parenrightbigg\ne±ik−\nSCxSk/bardbl(r/bardbl), (5)\nwhereu0andv0are the coherence factors expressed as\nu2\n0= 1−v2\n0=1\n2/bracketleftBig\n1+√\nE2−∆2\nE/bracketrightBig\n,andk+(−)\nSCis thexcom-\nponent of the wave number of an electron-(hole-)like QP\ndefined as k±\nSC=√\n2m\n/planckover2pi1/radicalBig\nµF±√\nE2−∆2−E/bardbl.\nThe wave function of the FM/NM/SC double junction\nis given by the linear combination of the above general\nsolutions. Let us consider the scattering of an electron\nin thek/bardblchannel with σ-spin injected into the NM from\nthe FM, the eight processesshown in Fig.1 (b) areactive.\nTherefore, the wave function in the FM layer ( x <0)\ntakes the form\nΨFM\nσ,k/bardbl(r) =/bracketleftBigg/parenleftbigg\n1\n0/parenrightbigg\neik+\nFM,σx+aσ,k/bardbl/parenleftbigg\n0\n1/parenrightbigg\neik−\nFM,σx\n+bσ,k/bardbl/parenleftbigg\n1\n0/parenrightbigg\ne−ik+\nFM,σx/bracketrightBigg\nSk/bardbl(r/bardbl). (6)In the NM layer (0 ≤x < d), we have\nΨNM\nσ,k/bardbl(r) =/bracketleftBigg\nασ,k/bardbl/parenleftbigg\n1\n0/parenrightbigg\neik+\nNMx+βσ,k/bardbl/parenleftbigg\n0\n1/parenrightbigg\ne−ik−\nNMx\n+ξσ,k/bardbl/parenleftbigg\n1\n0/parenrightbigg\ne−ik+\nNMx+χσ,k/bardbl/parenleftbigg\n0\n1/parenrightbigg\neik−\nNMx/bracketrightBigg\nSk/bardbl(r/bardbl),\n(7)\nand in the SC layer ( x≥d)\nΨSC\nσ,k/bardbl(r) =/bracketleftBigg\ncσ,k/bardbl/parenleftbigg\nu0\nv0/parenrightbigg\neik+\nSCx\n+dσ,k/bardbl/parenleftbigg\nv0\nu0/parenrightbigg\ne−ik−\nSCx/bracketrightBigg\nSk/bardbl(r/bardbl).(8)\nThe coefficients aσ,k/bardbl,bσ,k/bardbl,cσ,k/bardbl,dσ,k/bardbl,ασ,k/bardbl,βσ,k/bardbl,ξσ,k/bardbl,\nandχσ,k/bardblare determined by matching the wave func-\ntion at the boundary of the contact x= 0 and d. Fol-\nlowing the BTK theory19the probabilities of the AR\nand the normal reflection are given by Aσ,k/bardbl(E) =\n(k−\nFM,σ/k+\nFM,σ)a∗\nσ,k/bardblaσ,k/bardblandBσ,k/bardbl(E) =b∗\nσ,k/bardblbσ,k/bardbl, re-\nspectively. Since we assume that the temperature is\nzero, the conductance at bias voltage Vis given by\nG=e\nh/summationtext\nσ,k/bardbl/bracketleftbig\n1+Aσ,k/bardbl(eV)−Bσ,k/bardbl(eV)/bracketrightbig\n,whereweas-\nsume that the voltage drop occurs at the NM/SC inter-\nface for simplicity. Below the superconducting gap, i.e.,\neV <∆0, the probabilities Aσ,k/bardbl(E) andBσ,k/bardbl(E) sat-\nisfy the relation that 1 −Bσ,k/bardbl(E) =Aσ,k/bardbl(E) and then\nwe have\nG= 2e\nh/summationdisplay\nσ,k/bardblAσ,k/bardbl(eV). (9)\nLet us first consider the most idealistic case where the\ninterfacial scattering potential, V(x), is assumed to be\nzero. This simplification enables us to obtain the analyt-\nical expression of Aσ,k/bardbl(E) under the Andreev approxi-\nmation.\nAσ,k/bardbl(E)≃4(1−ζ)/radicalbig\n1−η2\n2/bracketleftBig\n(1−ζ)2+(1−ζ)/radicalbig\n(1−ζ)2−η2)/bracketrightBig\n−η2/bracketleftbig\nǫ2+(1−2ǫ2)cos2[(k+\nNM−k−\nNM)d]−ǫ√\n1−ǫ2sin[2(k+\nNM−k−\nNM)d]/bracketrightbig,\n(10)\nwhere we introduced the normalized parameters η=\nh0/µF,ζ=E/bardbl/µF, andǫ=E/∆0. We note that Eq.\n(10) contains trigonometric functions in the denomina-\ntor. For the FM/SC junction, i.e., d=0, the trigono-\nmetric functions become constant and Eq. (10) repro-\nduces the deJong and Beenakker results of the zero-bias\nconductance1. FortheFM/NM/SCjunctionswith d/negationslash= 0,thetrigonometricfunctionsinEq. (10)giverisetotheos-\ncillationofthe conductanceagainstthe biasvoltage. The\norigin of the conductance oscillation is the interference of\nelectrons in the NM layer11,15. In the FM/NM/SC dou-\nble junctions, the injected electron propagates across the\nNM layer to the interface as an electron (3 in Fig.1) and\nis scattered into a hole (6 in Fig.1) by the superconduct-3\nFIG. 2: (a) Probability of the AR Aσ,0(eV) for the\nFM/NM/SC double junction with d=1µm is plotted against\nthe bias voltage V. The exact numerical results and the value\nof Eq. (10) are plotted by lines and circles, respectively. T he\nvalue of the exchange field is taken to be h= 0.3, 0.6, 0.9 µF\nfrom top to bottom. (b) Same plot for d=10µm.\ning gap. The superconducting gap can pair an excited\nelectron with an electron inside the Fermi sea, leaving\na hole excitation. The hole propagates back across the\nNM layer; however, it cannot interfere with the original\nelectron. In order for interference to occur the hole must\nbe reflected at the FM/NM interface, propagate to the\nNM/SC interface, be scattered into the electron state (5\nin Fig.1) by the superconducting gap, and propagate as\nan electron in the NM layer. It can interfere with the\noriginal electron (3 in Fig.1). This interference produces\nan oscillation of the conductance against the bias volt-\nage, and the period of the oscillation is determined by\nthe thickness of the NM layer.\nIn order to analyze the interference effect on the con-\nductance oscillation, we consider the AR probability\nAσ,k/bardbl(E) of the one-dimensional system; i.e., only the\ntransverse channel with k/bardbl= 0 is considered. In Fig.\n2 (a), 2 (b) we plot the probability Aσ,k/bardbl(eV) of the\nFM/NM/SC double junctions with d=1µm and 10 µm,\nrespectively, as a function of the bias voltage. Since\nEq.(10) is an even function of normalized value of the\nexchange field η,Aσ,k/bardbl(eV) is independent of the spin di-\nrectionσ; i.e.,A↑,k/bardbl(eV) =A↓,k/bardbl(eV). TheFermienergy\nand the superconducting gap are assumed to be µF= 3.8\neV (kF= 1.0˚A−1) and ∆ 0= 1.5meV, respectively. The\nexact numerical results and the approximate values of\nEq. (10) are plotted by lines and circles, respectively.\nThe value of the exchange field is taken to be h0= 0.3,\n0.6, 0.9µFfrom top to bottom.\nAs shown in Figs. 2 (a) and 2 (b), Eq. (10) and\ntherefore the Andreev approximation are valid for all\nvalues of the exchange field. According to Eq. (10),\nthe period of the oscillation is determined by the condi-\ntion that k+−k−=nπ, wherenis an integer. Since\nk±≃kF(1±E/2µF), the period is obtained as\n∆V1D≃/planckover2pi12πkF\n2med, (11)\nwhich is inversely proportional to the thickness of the\nNM layer, d. For the one-dimensional system, the period\n∆V1DisindependentoftheexchangefieldoftheFMlayeras shown in Figs. 2 (a) and 2 (b), and in Eq.(11). How-\never,asweshallshowlater,theperiodoftheconductance\noscillation due to the geometrical resonance depends on\nthe exchange field of the FM layer because the number\nofk/bardblchannels available for the AR is restricted by the\nexchange field.\nTheperiodoftheoscillationof Aσ,k/bardbl(eV)withfinite k/bardbl\nis given by ∆ Vk/bardbl≃/planckover2pi1π√\n2med/radicalbigµF−E/bardbl. From Eq.(9) the\nconductanceisobtainedbysummingup Aσ,k/bardbl(eV) forall\navailable k/bardbl. Since the spin of the Andreev reflected hole\nis opposite to that of the incident electron, the maximum\nvalue of k/bardbland therefore E/bardblis limited by the exchange\nfieldh0as maxE/bardbl=µF−h0. We assume that oscilla-\ntions ofAσ,k/bardbl(eV) with different periods cancel out each\nother and the period of the sum/summationtext\nσ,k/bardblAσ,k/bardbl(eV) is de-\ntermined by the shortest period. Thus, the period of the\nconductance oscillation of the three-dimensional system\nis obtained as\n∆V3D≃min∆Vk/bardbl=/planckover2pi1π√\n2med/radicalbig\nh0.(12)\nIn Figs. 3 (a), 3 (b), and 3 (c) we plot the conductance\nof the FM/NM/SC junction, GFNS, normalized by that of\nthe FM/NM/NM junction, GFNN, against the bias volt-\nage. As shown in Fig. 3 (a), the oscillation due to the ge-\nometrical resonance does not appear in the conductance-\nvoltage curve if the thickness of the NM layer, d, is less\nthan or of the order of nm. The conductance-voltage\ncurve is indistinguishable from that of the FM/SC junc-\ntion. Hence, we can use the conventional PCAR analysis\nfor a FM film, the surface of which is coated by a thin\nFIG. 3: (a) The conductanceoftheFM/NM/SC double junc-\ntion,GFNSwithd=1nm is plotted against the bias voltage.\nThe conductance is normalized by that of the FM/NM/NM\njunction, GFNS. (b) Same plot for d=1µm. (c) Same plot for\nd=10µm. (d) The period of the conductance oscillation of the\nFM/NM/SC double junction with d=10µm is plotted against\nthe exchange field h0.4\nFIG. 4: (a) The normalized conductance, GFNS/GFNNfor\nd=1µm, Z=0.2 is plotted against the bias voltage. (b) Same\nplot ford=10µm, Z=0.2.\n(less than a few nm) NM layer.\nFigures 3 (b) and 3 (c) show the conductance-voltage\ncurves for the FM/NM/SC double junctions with d=\n1µm andd= 10µm, respectively. One can see that the\nperiod does depend on the exchange field, h0, in the FM\nlayer as well as the thickness of the NM layer. The pe-\nriod is a decreasing function of the exchange field. We\ncan easily confirm that the period is proportional to the\nsquare-root dependence of the exchange field by looking\nat Fig. 3 (d). The exact numerical results (filled cir-\ncles) agree well with Eq. (12) (dotted line). The results\nsuggest that we can determine the exchange field and\ntherefore the spin polarization of the FM layer from the\nperiod of the conductance oscillation.\nNext we consider the effect of the interfacial scattering\npotential at the NM/SC interface. We assume that the\ninterfacial potential is represented by the delta-function\nasV(x) = (/planckover2pi12kFZ/m)δ(x−d), where Zis the dimension\nless parameter which characterize the strength of the in-\nterfacial scattering potential. For simplicity, we neglect\nresistances of the FM/NM interface and the FM layer\nwhich do not change the oscillation period of the conduc-tance but reduce the amplitude of it. In Figs.4 (a) and 4\n(b), we show the normalized conductance-voltage curves\nfor junctions with d=1µm andd=10µm. The parameter\nZis assumed to be 0.26. Because the conductance os-\ncillation is due to the geometrical resonance in the NM\nlayer, the period of the oscillation is not affected by the\ninterfacial scattering potential and is given by Eq.(12).\nIn the present analysis we employed the simplest BdG\napproach and consider the clean FM/NM/SC junctions\nwith perfect interfaces. In the real experiments the con-\nductance oscillation we predicted might be smeared out\ndue to the interface roughness and imperfections. How-\never, recent advances in fabrication technology enables\nus to fabricate epitaxial FM/NM/SC trilayers of high\nquality22. We expect that the conductance oscillation\nwe predicted can be observed in such epitaxial trilay-\ners. For further understanding of the transport prop-\nerties of FM/NM/SC trilayers, we have to take into ac-\ncounttheeffectsoffinitemeanfreepath, bandstructures,\nand selfconsistent determination of the electron’s distri-\nbution function and electric-pottential, which is beyond\nthe scope of this Brief Report.\nIn summary, we studied the conductance oscillation\nduetothegeometricalresonanceinaFM/NM/SCdouble\njunction theoretically. We showed that the conductance\ndue to the Andreev reflection oscillates as a function of\nthe bias voltage due to the geometrical resonance. We\nfound that the exchange field and therefore the spin po-\nlarizationoftheFM layercanbedetermined fromthepe-\nriod of the conductance oscillation because the period of\nthe conductance oscillation is proportionalto the square-\nroot of the exchange field.\nThe authors would like to acknowledge the valuable\ndiscussions they had with K. Matsushita and N. Yokoshi.\n1M. J. M. de Jong and C. W. J. Beenakker, Phys. Rev. Lett.\n74, 1657 (1995).\n2R.J. Soulen Jr. et al., Science 282, 85 (1998).\n3S. K. Upadhyay, A. Palanisami, R. N. Louie, and R.\nA. Buhrman, Phys. Rev. Lett. 81, 3247 (1998).\n4R.J. Soulen Jr. et al., J. Appl. Phys. 85, 4589 (1999).\n5K. Kikuchi, H. Imamura, S. Takahashi, and S. Maekawa,\nPhys. Rev. B 65, 020508(R) (2001).\n6G. J. Strijkers, Y. Ji, F. Y. Yang, C. L. Chien, and J.\nM. Byers, Phys. Rev. B 63, 104510 (2001).\n7Y. Ji, G. J. Strijkers, F. Y. Yang, C. L. Chien, J. M. Byers,\nA. Anguelouch, G. Xiao, and A. Gupta, Phys. Rev. Lett.\n86, 5585 (2001).\n8I. I. Mazin, A. A. Golubov, and B. Nadgorny, J. Appl.\nPhys.89, 7576 (2001).\n9H. Imamura, K. Kikuchi, S. Takahashi, and S. Maekawa,\nJ. Appl. Phys. 91, 7032 (2002).\n10G. T. Woods et al., Phys. Rev. B 70, 054416 (2004).\n11P. G. de Gennes and D. Saint-James, Phys. Lett. 4, 151\n(1963).12W. J. Tomasch, Phys. Rev. Lett. 15, 672 (1965).\n13W. J. Tomasch, Phys. Rev. Lett. 16, 16 (1966).\n14W. L. McMillan and P. W. Anderson, Phys. Rev. Lett. 16,\n85 (1966).\n15J. M. Rowell and W. L. Macmillan, Phys. Rev. Lett. 16,\n453 (1966).\n16O. Nesher and G. Koren, Phys. Rev. B 60, 9287 (1999).\n17H.-S. Chang, M.-H. Bae, and H.-J. Lee, Physica C 408-\n410, 618 (2004).\n18L. Shkedy, P. Aronov, G. Koren, and E. Polturak, Phys.\nRev. B69, 132507 (2004).\n19G. E. Blonder, M. Tinkham, and T. M. Klapwijk, Phys.\nRev. B25, 4515 (1982).\n20P. G. de Gennes, Superconductivity of Metals and Alloys\n(W. A. Benjamin, New York, 1966), chap. 5.\n21L. R. Tagirov and N. Garc´ ıa, Superlattices Microstruct.\n41, 152 (2007).\n22H. Yamazaki, N. Shannon, and H. Takagi, Phys. Rev. B\n73, 094507 (2006)." }, { "title": "2105.09023v1.Determination_of_the_spin_Hall_angle_by_the_inverse_spin_Hall_effect__device_level_ferromagnetic_resonance__and_spin_torque_ferromagnetic_resonance__a_comparison_of_methods.pdf", "content": " 1 Determination of the spin Hall angle by the inverse spin Hall effect, device level ferromagnetic resonance, and spin torque ferromagnetic resonance: a comparison of methods Ranen Ben-Shalom1, Nirel Bernstein1, See-Hun Yang2, Amir Capua1* 1Department of Applied Physics, The Hebrew University of Jerusalem, Jerusalem 91904, Israel 2IBM Research Division, Almaden Research Center, 650 Harry Rd., San Jose, California 95120, USA *e-mail: amir.capua@mail.huji.ac.il Abstract: The spin torque ferromagnetic resonance (STFMR) is one of the popular methods for measurement of the spin Hall angle, 𝜽𝑺𝑯. However, in order to accurately determine 𝜽𝑺𝑯 from STFMR measurements, the acquired data must be carefully analyzed: The resonance linewidth should be determined to an accuracy of a fraction of an Oe, while the dynamical interaction leading to the measured response consists of the conventional field-induced ferromagnetic resonance (FMR), spin-torque induced FMR, and of the inverse spin Hall effect (ISHE). Additionally, the signal often deteriorates when DC current is passed through the device. In this work we compare the STFMR method with two other FMR-based methods that are used to extract 𝜽𝑺𝑯. The first is a device-level FMR and the second is based on the ISHE. We identify artefacts that are caused by the noise floor of the instrumentation that make the measurement of 𝜽𝑺𝑯 illusive even when the signal to noise ratio seems to be reasonable. Additionally, we estimate a 𝟏𝟎% error in 𝜽𝑺𝑯 that results from neglecting the magnetic anisotropies as in conventional measurements. Overall, we find the STFMR to be the most robust of the three methods despite the complexity of the interaction taking place therein. The conclusions of our work lead to a more accurate determination of 𝜽𝑺𝑯 and will assist in the search of novel materials for energy efficient spin-based applications. 2 Efficient generation of spin currents is key to spintronics. Over the last decade the majority of the efforts was focused on finding new physics for polarizing the transported spins. Such an example is the intensive study of the spin Hall effect (SHE) [1,2] that relies on the spin orbit coupling (SOC) or on non-trivial topologies of the band structure [3]. In addition to the challenges in generating spin currents, their accurate quantification is not straight forward. For example, in Pt, which is one of the main materials used to investigate the SHE, a large variation in the SHE efficiency was reported [4-7]. The charge to spin current conversion rate is referred to as the spin Hall angle [8], 𝜃(), and is defined by 𝜃()=2𝑒𝐽./ℏ𝐽1\twhere 𝑒 is the electron charge, ℏ is Planck’s constant, 𝐽1 is the charge current density, and 𝐽( is the spin current density. Measurement of 𝜃() can be carried out by various methods such as the cavity ferromagnetic resonance (FMR) [9,10], spin pumping [11-15], optical FMR [16,17], anomalous and planar Hall effect [18-21], spin torque ferromagnetic resonance (STFMR) [5,22-25], and nonlocal spin transport [8,26-28]. Among these methods, the STFMR is most commonly used due to its simplicity and applicability. In the STFMR, an AC charge current is injected into a bilayer of a ferromagnet (FM)/heavy metal (HM) and excites spin precessions. Due to the anisotropic magneto resistance (AMR) [29], a rectified DC voltage builds up. Accordingly, the measured FMR response stems from three different processes: 1) SHE in the HM that generates AC spin currents that are responsible for an antisymmetric lineshape. 2) A symmetric lineshape that stems from the stray AC Oersted field generated by the current passing in the HM. 3) An additional symmetric resonance response that originates in spin pumping [30] by the precessing magnetization 𝑀44⃗ [22] in the FM layer. Therefore, analysis of the measured lineshape is not trivial and requires separating the different contributions, each resolved to a fraction of an Oe. This, in turn, usually requires solving a five-parameter lineshape fitting optimization problem [24]. In this work, we compare the STFMR technique with the inverse SHE (ISHE) [31] and device-level ferromagnetic resonance (DLFMR) methods for measurement of 𝜃(). All techniques are applied on the same device. Each method has relative advantages such as the elimination of the spin transfer torque (STT) [22] or in the simplicity of the analysis of the measured responses. Our study shows that the anisotropies that are usually neglected, e.g. the shape anisotropy, may play a considerable role [32]. However, we show that the conventional optimization algorithm used to analyze the STFMR response is robust. In addition, our measurements indicate a significant electrical background noise that affects all methods. This noise often deteriorates the ISHE and DLFMR signals and in the worst case leads to unnoticeable measurement artefacts that result in erroneous values of 𝜃(). We conclude that despite the relative advantages of the ISHE and DLFMR methods and the challenge of analyzing the STFMR response the STFMR is a more robust and reliable technique. 3 Fig. 1. Experimental setup. (a) Patterned SHE device with a micro-antenna. (b) Schematic of the measurement setup that includes the STFMR, ISHE, and DLFMR. The experimental system is presented in Fig. 1 and can be converted between the STFMR, ISHE, and DLFMR modes of operation so that all methods are applied on the same device. In the STFMR a microwave signal is generated by the RF signal generator and is driven into the device by connecting the RF signal generator to node ‘A’ indicated in the figure. The injected AC current drives the FMR in the device in the presence of an externally applied DC magnetic field,\t𝐻789. The oscillatory driving torque stems from the Oersted stray field and the AC STT that is generated by the SHE. Due to the AMR and the magnetization precessions, the AC current is rectified and a DC voltage builds up on the device. In addition, the precessions of 𝑀44⃗ pump spin current by the spin pumping mechanism [30] into the HM which is converted into a DC charge current by the ISHE. Once the FMR is excited a DC charge current, 𝐼;, that is converted into a DC spin current is injected and exerts an anti-damping torque that is reflected on the resonance linewidth from which 𝜃() is extracted in the usual manner [22]. The DC voltage building upon the device in the STFMR configuration is given by 𝑉(=>?@=𝐼;𝑅;+𝐼C()D(𝐻789)⋅𝑅;+HI⋅𝑖;⋅(∆𝑅L7(𝐻789)+∆𝑅(==(𝐻789)) where 𝑖;, 𝑅;, 𝐼C()D, ∆𝑅L7, and ∆𝑅(== are the amplitude of AC charge current, the DC resistance, the DC ISHE charge current that is pumped by 𝑀44⃗, the amplitude of the AC resistance that originates in the AC Oersted excitation, and the amplitude of the AC resistance that originates from the STT, respectively. Both ∆𝑅L7 and ∆𝑅(== stem from the AMR. To extract the resonance linewidth, following a data fitting procedure, the measured resonances are decomposed to the symmetric and antisymmetric components that originate in the Oersted and STT excitations. The fitting process requires solving a five-parameter optimization problem of the following expression: \nAntenna Device \nFigure captions a b \n 4 𝑉(=>?@=𝐴∆)N()OPQR)SOT)NU∆)N+𝐵∆)N()OPQR)SOT)NU∆)N)OPQR)SOT∆)\t+𝐶 (1) In Eq. 1 𝐻X7Y is the resonance field, ∆𝐻 is the resonance width, and A, B, and C are the coefficients of the symmetric, antisymmetric, and DC background level, respectively. The contribution of the ISHE in 𝑉(=>?@ is neglected in our analysis [22]. Our experimental results below indicate that this approximation is valid, although according to Ref. [33] the ISHE contribution was found to be significant. In the other two configurations of the DLFMR and ISHE, the AC current is driven into a micro-antenna that is patterned in proximity to the device (Fig. 1) rather than being injected into it. In this manner we excite the FMR without the contribution of the AC STT. Therefore, the dynamical processes taking place in the ISHE and the DLFMR measurements are greatly simplified as compared to the STFMR. Consequently, the resonance responses of the ISHE and DLFMR are purely symmetric and in the data fitting procedure the optimization algorithm is applied to solve a four-parameter problem. The ISHE and DLFMR differ in the manner at which the FMR signal is read. In the ISHE spin pumping induces a DC voltage signal by the ISHE in the HM layer that is probed. The measured voltage is given by 𝑉C()D=𝐼;𝑅;+𝐼C()D(𝐻789)⋅𝑅;. In the DLFMR configuration the FMR signal is read at the fundamental RF harmonic. Due to the AMR effect, the resistance of the device is modulated at the RF frequency and in the presence of the applied DC current a voltage signal is generated at the fundamental frequency. A homodyne detection scheme using an RF mixer is then used to down-convert this signal and read it as illustrated in Fig. 1. The measured DLFMR signal is expressed by 𝑉Z[>?@=\\𝐼;+𝐼C()D(𝐻789)]⋅∆𝑅L7(𝐻789). The DLFMR signal contains also an ISHE contribution, 𝐼C()D, however, it is three orders of magnitude weaker than 𝐼;. To further eliminate ISHE component, in the experiment we used a lock-in detection scheme in which 𝐼; was modulated. In order to model the dynamics with a minimal set of approximations we derived the theoretical model based on the Smit-Beljiers-Suhl formalism [17,34,35]. Accordingly, the Landau-Lifshitz-Gilbert equation in spherical coordinates becomes: 𝜃̇=𝛾𝐻`,\t\t\t\t𝑠𝑖𝑛𝜃𝜙̇=−𝛾𝐻f (2) Here, 𝛾 is the gyromagnetic ratio while 𝐻f and 𝐻` are the effective fields along the polar and azimuthal unit vectors, respectively. These originate from the conserving energies, 𝐸hijY7Xk., that account for 𝐻789, the demagnetizing field, the crystalline anisotropy, the microwave field, and the field-like spin orbit torque (SOT):\t𝐸hijY7Xk.=−𝑀44⃗⋅\\𝐻44⃗\t789+𝐻44⃗L7+ℎ4⃗L7]−12𝑀44⃗⋅𝐻44⃗Z−𝐾psinI𝜃−𝑀44⃗⋅ℎ4⃗Xt−𝐻>[𝑀44⃗⋅𝑠̂. 𝐻44⃗L7, ℎ4⃗L7, 5 𝐻44⃗Z, ℎ4⃗Xt, 𝐻44⃗>[, 𝑠̂, and 𝐾p are the stray DC Oersted field stemming from the HM, the Oersted field generated by AC current, the demagnetization field, the RF field of amplitude ℎXt, the field-like STT, the spin polarization of the injected spins, and the magnetocrystalline anisotropy constant, respectively. In addition, 𝐻f and 𝐻` also stem from the nonconserving energies, 𝐸jijhijY7Xk. that account for the magnetic Gilbert damping losses and the damping-like SOT:\t𝐸jijhijY7Xk.=(𝛼/2𝛾𝑀Y)𝑀44⃗̇I+𝐻()D𝑀44⃗̇⋅\\𝑀44⃗×𝑠̂] where 𝛼 is the Gilbert damping and 𝑀Y is the saturation magnetization. 𝐻f and 𝐻` are then determined from the effective field, 𝐻44⃗7tt, that is given by 𝐻44⃗7tt=−∇?44⃗𝐸hijY7Xk.−∇?44⃗̇𝐸jijhijY7Xk.. The spin Hall parameter 𝐻()D is given by ℏ𝜃()𝐽1/2𝑒𝑀(𝑡\twhere 𝑡 is the thickness of the ferromagnetic layer. 𝐸hijY7Xk. and 𝐸jijhijY7Xk. have units of energy density and energy density flow. Solving for small deviations, (∆𝜃,∆𝜙) around the equilibrium state, (𝜃;,𝜙;), in phasor-space we find the in-plane component of 𝑀44⃗ at the angular frequency 𝜔 which is given by: ∆𝜙={(HU|N)}~f\\NRN]U∆\\NRN]NUN∆N\\ℎf+𝛼ℎ`]𝑖𝜔+{?TH}~fℎf𝜁`f+ℎ`𝜁ff (3) with: 𝜔;I=𝛾I𝑀YI(1+𝛼I)sinI𝜃;𝜁``𝜁ff−𝜁`f𝜁f` ∆𝜔=𝛾𝑀Y(1+𝛼I)sin𝜃;𝛼sin𝜃;𝜁ff+𝛼sin𝜃;𝜁``+𝜁`f−𝜁f` And 𝜁 defined by 𝜁f`=𝐸f`,𝜁`f=𝐸`f+𝑀Y𝐻()D\\𝑠8cos𝜙;−𝑠sin𝜙;] 𝜁ff=𝐸ff+𝑀Y𝐻()Dsin𝜃;\\−𝑠8cos𝜃;sin𝜙;+𝑠cos𝜃;cos𝜙;] 𝜁``=𝐸``+𝑀Y𝐻()Dsin𝜃;\\𝑠8sin𝜃;cos𝜙;+𝑠sin𝜃;sin𝜙;−𝑠cos𝜃;] In Eq. (3) 𝜔; is the resonance frequency, ∆𝜔 is the resonance linewidth, and ℎf and ℎ` are the polar and azimuthal components of the RF excitation. 𝑠8,\t𝑠, and 𝑠\tare the projection of 𝑠̂ on the cartesian axis. 𝐸``,𝐸`f,𝐸f`, and 𝐸ff are the second order derivatives of (𝐸hijY7Xk.+𝐸jijhijY7Xk.), with respect to 𝜃 and 𝜙. The equilibrium state was evaluated without the contribution of the SOT. 6 To compare the methods we used the well-studied Pt based system: 60 Pt/5 Py/60 Co/20 TaN (units in Å). The film was grown by magnetron sputtering and patterned to a 5\t𝑋\t40\t𝜇𝑚 device. The Au micro-antenna of 100\t𝑛𝑚 thickness and 2\t𝜇𝑚 width was deposited at a distance of 1.5\t𝜇𝑚. The AC current density driven into the device in the STFMR experiment was 10⋅10HI\t𝐴/𝑚I and 1.4⋅10HI\t𝐴/𝑚I into micro-antenna in the ISHE and DLFMR experiments. To maximize the signals, in the STFMR and DLFMR experiments the current was injected at 45° with respect to 𝐻789 and at 90° in the ISHE experiment. The frequency dispersion relation and linewidth vs. frequency as obtained from the STFMR method are presented in Fig. 2. From these measurements the magnetic parameters of the sample were extracted using Kittel’s formula, 𝑓;=𝛾/2𝜋⋅𝐻X7Y(𝐻X7Y+4𝜋𝑀Y), with 𝑓; being the resonance frequency, and the frequency dependent linewidth, ∆𝐻=2α/γ⋅𝜔+∆𝐻C), with ∆𝐻C) being the inhomogeneous broadening. Accordingly, the values of 𝛼=0.0048,\t𝑀Y=1384\t𝑒𝑚𝑢/𝑐𝑚£, and ∆𝐻C)=16\t𝑂𝑒 were extracted. \n Fig. 2. (a) STFMR Frequency-field dispersion relation. Open blue circles indicate the measurement. Solid red line indicates Kittel’s formula. Inset illustrates the measured STFMR responses for 5 - 18 GHz. (b) Resonance linewidth as a function of frequency. Solid red line indicates the linear fit. The measurement of 𝜃() using the STFMR and ISHE experiments are presented in Fig. 3. The experiments were carried out at 8 GHz. The figure presents ∆𝐻 and 𝐻X7Y as function of 𝐽1. Fig. 3(a) and 3(b) present the STFMR results. The shift of 𝐻X7Y with 𝐽1 originates from the stray DC Oersted field stemming from the DC current and agrees well with a calculation of the expected stray field. The linewidth measurements were fitted using Eq. (3) from which a 𝜃() of 23.1% was extracted. This value is within the range of previously measured 𝜃() values [4,17,36,37]. Interestingly, it is readily seen that the noise level increases as |𝐽1| is a b 5GHz 18GHz 7 increased. This behavior is marked by the shaded area in the figure and is typical of an intrinsic Johnson noise and the shot noise limited processes. These processes take place either in the device or the measurement instrumentation, or both. This noise limits the maximal DC current that can be injected. A similar behavior was also seen in Ref. [38]. There, it was shown that this limitation can be overcome by optically probing 𝑀44⃗. \n Fig. 3. (a) 𝑯𝒓𝒆𝒔 as a function of 𝑱𝑪 for the STFMR measurements. (b) ∆𝑯 as a function of 𝑱𝑪 for the STFMR measurements. Shaded blue area guides the eye to the noise level. Black dashed line indicates ∆𝑯 at 𝑱𝑪 = 0. (c) 𝑯𝒓𝒆𝒔 as a function of 𝑱𝑪 for the ISHE measurements. (d) ∆𝑯 as a function of 𝑱𝑪 for the ISHE measurements. Black dashed line marks ∆𝑯 for 𝑱𝑪 = 0. In (a) - (d) the open blue circles indicate the measurement and solid red lines indicate the linear fits. Data presented at 8 GHz. STFMR \nISHE ISHE STFMR a b \nc d \n 8 The overall broadening that results from the anti-damping torque is only of a few Oe. Therefore, in order to extract 𝜃(), the linewidth should be determined to the accuracy of a fraction of an Oe. However, since the optimization algorithm is applied on a multi-parameter problem, five in our case, it is possible that it would converge into a local minimum. Therefore, we tested the convergence of the algorithm. This was carried out by providing the algorithm with initial conditions that are far off from the actual solution. We found that a shift larger than 100 Oe in the resonance frequency and 50 Oe in the linewidth was the limit after which the optimization algorithm converged to a wrong solution. This illustrates that the fitting algorithm is robust. The dependence of 𝐻X7Y and ∆𝐻 on the injected DC current in the ISHE measurement are shown in Fig. 3(c) and 3(d). The ISHE measurements are significantly noisier. Consequently, a wrong value of 𝜃()=46.8 that is larger by a factor of two as compared to the STFMR method was measured. Despite the noise, we can readily see that the shift in 𝐻X7Y as function of 𝐽1 is much smaller than in the STFMR measurement. In the STFMR the overall shift is ~ 35 Oe while in the ISHE the shift is only ~ 11 Oe. Since the STFMR measurements are carried out at 45° and the ISHE measurements at 90° we can attribute the difference to the anisotropy fields which are usually neglected. However, our model which includes the anisotropy fields could not fully account for the differences and we believe there is an additional contribution such as the existence of magnetic domains. From our model, we found that neglecting the anisotropy fields results in an error of ~10% in 𝜃(). The DLFMR configuration is the most direct measurement of the FMR. The FMR is excited externally by the micro-antenna and is read directly on the device using a homodyne detection scheme. In contrast to the other configurations, in the DLFMR the signal arises only from the AC resistance that is modulated by the AMR. Additionally, the homodyne detection scheme eliminates the DC noise which is responsible for a significant portion of the noise as shown above. For these reasons we expect the DLFMR experiment to be more accurate. In reality this is not the case and the DLFMR measurement (Fig. 4(a)) was noisy to the extent that we could not extract 𝜃() reliably. We speculate that the noise originates from thermal effects in the RF mixer. In order to compare the STFMR with the ISHE and DLFMR methods we plot only the symmetric part of the STFMR measurement that originates in the Oersted field. The symmetric part of the STFMR response has the same functional form as the DLFMR and ISHE lineshapes. A comparison of all lineshapes overlaid on top of each other is presented in Fig. 4. Each trace was normalized to its maximum. The ISHE resonance frequency is higher by ~ 50 Oe as compared to the STFMR, and DLFMR. This stems from the difference 9 in the direction at which 𝐼; is applied with respect to 𝐻789. This behavior demonstrates once more that the anisotropies are not as negligible as conceived. The shift in the resonance field resulting from the different angle of applied 𝐻789 is also seen in the STFMR and ISHE measurements of 𝐻X7Y as function of 𝐽1\tin Fig. 3. Fig. 4(a) shows that the DLFMR and ISHE are much noisier as compared to the STFMR. At the origin of the different noise levels of the ISHE and the DLFMR lies the physical mechanism used to read out the signal. In the ISHE, the signal originates from the spin pumping and in the DLFMR and STFMR it originates in the AMR. These differences translate to a STFMR signal that is 26 times stronger than the ISHE signal and, therefore, lead to a higher signal to noise ratio (SNR). For this reason, conclude that the contribution of the ISHE is negligible in the STFMR experiment and can be neglected. In the DLFMR the signal originates in the AMR as in the STFMR. Hence, the magnitude of the signals readout is similar. In the DLFMR, the absolute noise level is higher than in the STFMR because of the mixer and possibly impedance mismatching effects that are not present in the DC measurements. Therefore, the SNR is significantly lower than in the STFMR measurement. To understand how the measurement is affected by the poor SNR we simulated a Lorentzian resonance function that is submerged in a background noise level as shown in Fig. 4(b). We then fitted the resultant signal to an ideal Lorentzian. It is seen that the fitted resonance is narrower than the real resonance. The narrower linewidth of the ISHE and DLFMR is also seen in our measurements when the ISHE lineshape is overlaid on the STFMR and DLFMR lineshapes as illustrated by the black dashed line in Fig. 4(a). Clearly, the DLFMR and ISHE resonances appear narrower than the STFMR resonance indicating the significant contribution of the noise. We evaluate the ISHE noise floor to be 0.8𝜇𝑉 and the DLFMR noise floor to be 140𝜇𝑉. Both are 7% of the resonance peak. We can see the same effect in the linewidth measurements of the STFMR and ISHE in Fig. 3. The absolute width of the ISHE spectrum is smaller than the STFMR width as indicated by the vertical black dashed lines of Fig. 3(b) and 3(d). This illusive artifact illustrates the challenges of reliably extracting 𝜃(). 10 Fig. 4. (a) Symmetric part of the STFMR measurement (blue) together with the responses of the DLFMR (red) and ISHE (yellow) measurements and the shifted ISHE response (dashed black). Traces are normalized to the peak value. (b) Noise simulation. Blue solid line indicates an ideal Lorentzian lineshape. Red solid line illustrates the calculated Lorentzian lineshape together with a simulated noise floor. Black solid line is a fit of the simulated noisy Lorentzian to a noise-free response. To summarize, in this work we have examined the intricate details of the commonly used STFMR technique and compared it to the ISHE and DLFMR methods that sense more directly the FMR response in the presence of spin currents. We found that despite the physical complexity of the dynamical interaction in the STFMR, it is more robust than the ISHE and DLFMR. We showed that it is necessary to consider the anisotropies, e.g. the shape anisotropy in the device to resolve more accurately 𝜃(). Our study shows that the background noise distorts the measured signal even when the SNR is high enough, resulting in erroneous 𝜃() values and may be one of the reasons for the large scattering in the values reported in the literature for common materials. Although the DLFMR is expected to be the most reliable method, there seems to be a fundamental limitation to the measurement. Our work marks another step towards utilizing the SHE in realistic practical applications. Lorentz. line Lorentz. + noise a b STFMR symmetric part ISHE DLFMR fitShifted ISHE 11 References [1] J. E. Hirsch, \"Spin Hall Effect\", Physical Review Letters 83, 1834 (1999). [2] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, \"Spin Hall effects\", Reviews of Modern Physics 87, 1213 (2015). [3] Y. Zhang, J. Železný, Y. Sun, J. van den Brink, and B. Yan, \"Spin Hall effect emerging from a noncollinear magnetic lattice without spin–orbit coupling\", New Journal of Physics 20, 073028 (2018). [4] E. Sagasta, Y. Omori, M. Isasa, M. Gradhand, L. E. Hueso, Y. Niimi, Y. Otani, and F. Casanova, \"Tuning the spin Hall effect of Pt from the moderately dirty to the superclean regime\", Physical Review B 94, 060412 (2016). [5] K. Kondou, H. Sukegawa, S. Mitani, K. Tsukagoshi, and S. Kasai, \"Evaluation of Spin Hall Angle and Spin Diffusion Length by Using Spin Current-Induced Ferromagnetic Resonance\", Applied Physics Express 5, 073002 (2012). [6] Z. Feng, J. Hu, L. Sun, B. You, D. Wu, J. Du, W. Zhang, A. Hu, Y. Yang, D. M. Tang, B. S. Zhang, and H. F. Ding, \"Spin Hall angle quantification from spin pumping and microwave photoresistance\", Physical Review B 85, 214423 (2012). [7] W. Zhang, V. Vlaminck, J. E. Pearson, R. Divan, S. D. Bader, and A. Hoffmann, \"Determination of the Pt spin diffusion length by spin-pumping and spin Hall effect\", Applied Physics Letters 103, 242414 (2013). [8] S. O. Valenzuela and M. Tinkham, \"Direct electronic measurement of the spin Hall effect\", Nature 442, 176 (2006). [9] K. Ando, S. Takahashi, K. Harii, K. Sasage, J. Ieda, S. Maekawa, and E. Saitoh, \"Electric Manipulation of Spin Relaxation Using the Spin Hall Effect\", Physical Review Letters 101, 036601 (2008). [10] S. Emori, T. Nan, T. M. Oxholm, C. T. Boone, J. G. Jones, B. M. Howe, G. J. Brown, D. E. Budil, and N. X. Sun, \"Quantification of the spin-Hall anti-damping torque with a resonance spectrometer\", Applied Physics Letters 106, 022406 (2015). [11] O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bader, and A. Hoffmann, \"Quantifying Spin Hall Angles from Spin Pumping: Experiments and Theory\", Physical Review Letters 104, 046601 (2010). [12] V. Vlaminck, J. E. Pearson, S. D. Bader, and A. Hoffmann, \"Dependence of spin-pumping spin Hall effect measurements on layer thicknesses and stacking order\", Physical Review B 88, 064414 (2013). [13] F. D. Czeschka, L. Dreher, M. S. Brandt, M. Weiler, M. Althammer, I. M. Imort, G. Reiss, A. Thomas, W. Schoch, W. Limmer, H. Huebl, R. Gross, and S. T. B. Goennenwein, \"Scaling Behavior of the Spin Pumping Effect in Ferromagnet-Platinum Bilayers\", Physical Review Letters 107, 046601 (2011). [14] L. Bai, P. Hyde, Y. S. Gui, C. M. Hu, V. Vlaminck, J. E. Pearson, S. D. Bader, and A. Hoffmann, \"Universal Method for Separating Spin Pumping from Spin Rectification Voltage of Ferromagnetic Resonance\", Physical Review Letters 111, 217602 (2013). [15] D. Wei, M. Obstbaum, M. Ribow, C. H. Back, and G. Woltersdorf, \"Spin Hall voltages from a.c. and d.c. spin currents\", Nature Communications 5, 3768 (2014). [16] A. Capua, S.-H. Yang, T. Phung, and S. S. P. Parkin, \"Determination of intrinsic damping of perpendicularly magnetized ultrathin films from time-resolved precessional magnetization measurements\", Physical Review B 92, 224402 (2015). [17] A. Capua, T. Wang, S.-H. Yang, C. Rettner, T. Phung, and S. S. P. Parkin, \"Phase-resolved detection of the spin Hall angle by optical ferromagnetic resonance in perpendicularly magnetized thin films\", Physical Review B 95, 064401 (2017). 12 [18] K.-F. Huang, D.-S. Wang, H.-H. Lin, and C.-H. Lai, \"Engineering spin-orbit torque in Co/Pt multilayers with perpendicular magnetic anisotropy\", Applied Physics Letters 107, 232407 (2015). [19] K. Garello, I. M. Miron, C. O. Avci, F. Freimuth, Y. Mokrousov, S. Blugel, S. Auffret, O. Boulle, G. Gaudin, and P. Gambardella, \"Symmetry and magnitude of spin-orbit torques in ferromagnetic heterostructures\", Nat Nano 8, 587 (2013). [20] U. H. Pi, K. Won Kim, J. Y. Bae, S. C. Lee, Y. J. Cho, K. S. Kim, and S. Seo, \"Tilting of the spin orientation induced by Rashba effect in ferromagnetic metal layer\", Applied Physics Letters 97, 162507 (2010). [21] J. Kim, J. Sinha, M. Hayashi, M. Yamanouchi, S. Fukami, T. Suzuki, S. Mitani, and H. Ohno, \"Layer thickness dependence of the current-induced effective field vector in Ta|CoFeB|MgO\", Nat Mater 12, 240 (2013). [22] L. Liu, T. Moriyama, D. C. Ralph, and R. A. Buhrman, \"Spin-Torque Ferromagnetic Resonance Induced by the Spin Hall Effect\", Physical Review Letters 106, 036601 (2011). [23] A. Azevedo, L. H. Vilela-Leão, R. L. Rodríguez-Suárez, A. F. Lacerda Santos, and S. M. Rezende, \"Spin pumping and anisotropic magnetoresistance voltages in magnetic bilayers: Theory and experiment\", Physical Review B 83, 144402 (2011). [24] K.-U. Demasius, T. Phung, W. Zhang, B. P. Hughes, S.-H. Yang, A. Kellock, W. Han, A. Pushp, and S. S. P. Parkin, \"Enhanced spin–orbit torques by oxygen incorporation in tungsten films\", Nature Communications 7, 10644 (2016). [25] K. Kondou, H. Sukegawa, S. Kasai, S. Mitani, Y. Niimi, and Y. Otani, \"Influence of inverse spin Hall effect in spin-torque ferromagnetic resonance measurements\", Applied Physics Express 9, 023002 (2016). [26] C. Brune, A. Roth, E. G. Novik, M. Konig, H. Buhmann, E. M. Hankiewicz, W. Hanke, J. Sinova, and L. W. Molenkamp, \"Evidence for the ballistic intrinsic spin Hall effect in HgTe nanostructures\", Nat Phys 6, 448 (2010). [27] K. Olejník, J. Wunderlich, A. C. Irvine, R. P. Campion, V. P. Amin, J. Sinova, and T. Jungwirth, \"Detection of Electrically Modulated Inverse Spin Hall Effect in an $\\mathrm{Fe}/\\mathrm{GaAs}$ Microdevice\", Physical Review Letters 109, 076601 (2012). [28] L. Vila, T. Kimura, and Y. Otani, \"Evolution of the Spin Hall Effect in Pt Nanowires: Size and Temperature Effects\", Physical Review Letters 99, 226604 (2007). [29] T. R. McGuire and R. I. Potter, \"Anisotropic Magnetoresistance in Ferromagnetic 3d Alloys\", IEEE Transactions on Magnetics 11, 1018 (1975). [30] Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, \"Enhanced Gilbert Damping in Thin Ferromagnetic Films\", Physical Review Letters 88, 117601 (2002). [31] E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, \"Conversion of spin current into charge current at room temperature: Inverse spin-Hall effect\", Applied Physics Letters 88 (2006). [32] W. Zhang, W. Han, X. Jiang, S.-H. Yang, and S. S. P. Parkin, \"Role of transparency of platinum-ferromagnet interfaces in determining the intrinsic magnitude of the spin Hall effect\", Nature Physics 11, 496 (2015). [33] Y. Zhang, Q. Liu, B. F. Miao, H. F. Ding, and X. R. Wang, \"Anatomy of electrical signals and dc-voltage line shape in spin-torque ferromagnetic resonance\", Physical Review B 99, 064424 (2019). [34] J. Smit and H. G. Beljers, \"Ferromagnetic resonance absorption in BaFe12O19, a highly anisotropic crystal\", Philips Research Reports 10, 113 (1955). [35] H. Suhl, \"Ferromagnetic Resonance in Nickel Ferrite Between One and Two Kilomegacycles\", Physical Review 97, 555 (1955). [36] T. Kimura, Y. Otani, T. Sato, S. Takahashi, and S. Maekawa, \"Room-Temperature Reversible Spin Hall Effect\", Physical Review Letters 98, 156601 (2007). 13 [37] T. Seki, Y. Hasegawa, S. Mitani, S. Takahashi, H. Imamura, S. Maekawa, J. Nitta, and K. Takanashi, \"Giant spin Hall effect in perpendicularly spin-polarized FePt/Au devices\", Nature Materials 7, 125 (2008). [38] B. Grover, B. Hazara, B. Pal, T. Ma, S. Choudhary, J.-C. Jeon, N. Bernstein, A. Capua, and S. Parkin, in American Physics Society (APS), March meeting 2021 USA, 2021). " }, { "title": "1403.0656v1.Off_Resonant_Manipulation_of_Spins_in_Diamond_via_Precessing_Magnetization_of_a_Proximal_Ferromagnet.pdf", "content": "O\u000b-Resonant Manipulation of Spins in Diamond via Precessing Magnetization of a\nProximal Ferromagnet\nC. S. Wolfe*,1V. P. Bhallamudi*,1H. L. Wang,1C. H. Du,1S.\nManuilov,1A. J. Berger,1R. Adur,1F. Y. Yang,1and P. C. Hammel1,\u0003\n1Department of Physics, The Ohio State University, Columbus, Ohio 43210, USA\n(Dated: September 29, 2018)\nWe report the manipulation of nitrogen vacancy (NV) spins in diamond when nearby ferrimagnetic\ninsulator, yttrium iron garnet, is driven into precession. The change in NV spin polarization, as\nmeasured by changes in photoluminescence, is comparable in magnitude to that from conventional\noptically detected magnetic resonance, but relies on a distinct mechanism as it occurs at a microwave\nfrequency far removed from the magnetic resonance frequency of the NV spin. This observation\npresents a new approach to transferring ferromagnetic spin information into a paramagnet and\nthen transducing the response into a robust optical signal. It also opens new avenues for studying\nferromagnetism and spin transport at the nanoscale.\nPACS numbers: 72.25.Mk, 75.76.+j, 75.78.-n, 75.78.-n, 75.78.-n\nKeywords: NV, magnetization dynamics, FMR, spin transport, YIG\nUnderstanding the transport of spin and energy be-\ntween dissimilar materials is a topic of intense current\ninterest re\recting both the scienti\fc richness of the topic\nas well as it technological potential[1{6]. Metal/metal in-\nterfaces have been extensively studied, and to a lesser de-\ngree metal/semiconductor and metal/insulator systems.\nHowever, the transfer of angular momentum between two\ninsulating materials has been more challenging to study\ndue to the lack of suitable detection methods.\nNV centers in wide band-gap insulating diamond pro-\nvide an exceptional platform for performing spin-based\nmeasurements. The paramagnetic NV center is optically\nactive, and its photoluminescence (PL) is dependent on\nthe relative occupation of the lowest lying electronic spin\nstate of the defect center [7, 8]. This enables optical mea-\nsurement of NV-center spin state with excellent sensitiv-\nity, making optically detected magnetic resonance of NV\ncenters an area of intense research activity [9{21]. How-\never, work done thus far relies on manipulation of NV\nspins using magnetic resonance.\nHere we present experimental evidence that the NV\ncenter state can be modi\fed non-resonantly ( i.e.,by ir-\nradiation with microwave magnetic \felds at frequencies\nfar from NV center Larmor frequencies) by coupling to\nthe dynamics of a proximal ferromagnetic insulator. This\nchange can be detected as a change in the NV center\nPL, as is done for conventional microwave-driven reso-\nnant spin manipulations. This new e\u000bect promises valu-\nable insights into interactions between spins in adjacent\ndissimilar materials. This is a particular case in which\nboth materials are insulating and the interaction is ef-\nfective over long distances ( >300 nm). More generally,\n* C. Wolfe and V. Bhallamudi contributed equally to this work.\nFIG. 1. Experimental Schematic : The sample is a 20 nm\nthick single crystal YIG \flm with nanodiamonds dispersed on\ntop with a thickness of about 500 nm. To apply microwave\n\felds to the sample a silver microwire is patterned on the YIG.\nGreen laser light is focused onto nanodiamonds near the wire,\nand the intensity of the resulting photoluminescence from the\nNV centers is measured. Inset is an SEM image of dispersed\nnanodiamonds.\nthis provides a novel method for manipulating NV center\nspins and could enable sensitive spatially resolved imag-\ning of ferromagnetic phenomena by means atomic scale\nNV centers [9, 12, 15].\nYttrium Iron Garnet (YIG), Y 3Fe5O12, was chosen for\nthis experiment as a well known ferrimagnetic insulator\nwith exceptionally low damping [22, 23]. Here an epitax-\nial YIG \flm, 20nm thick, was grown on a gadolinium\ngallium garnet (GGG) (111) substrate by o\u000b-axis sput-\ntering [24{26]. Continuous wave microwave \felds are ap-\nplied to the sample by means of a 300 nm thick and 30 \u0016m\nwide silver microstrip line patterned on top of the YIG.arXiv:1403.0656v1 [cond-mat.mes-hall] 4 Mar 20142\nNanodiamonds, 50-200 nm in size (as shown by SEM im-\nage analysis) and containing up to a few thousand NV\ncenters each, were dispersed on top of a lithographically\nde\fned microstrip line as shown in Fig. 1. AFM mea-\nsurement indicates that the nanodiamonds form a 500\nnm thick \flm. Photoluminescence is excited in the NV\ncenters using a 532 nm laser beam and is collected by a\nphotodiode. A lock-in measurement is performed on the\nphotodiode signal by modulating the amplitude of the\napplied microwave \feld.\nThe lock-in measurement of the resulting modulation\nof the PL intensity is presented in Fig. 2 as a function\nof an applied in-plane magnetic \feld and the applied mi-\ncrowave frequency. Data for a control sample with nan-\nodiamonds on a GGG substrate without YIG is shown\nin Fig. 2 (a). We observe the intrinsic and well-known\nmagnetic resonances of the NV center ground and excited\nstates, starting at 2.87 GHz and 1.43 GHz respectively.\nShown in Fig. 2 (b) is the same data overlaid with the\ntheoretically expected resonance conditions for the NV\ncenters, which are obtained by solving the NV hamilto-\nnian in the presence of magnetic \feld parallel and per-\npendicular to the NV axis (see supplementary informa-\ntion of [9]). We also see several features in the PL below\n1.25 GHz (Fig. 2 (a)) that are not related to the normal\nground- and excited-state resonances of the NV centers.\nThese will be discussed in a forthcoming publication.\nThe data obtained from the nanodiamonds on top of\nthe YIG \flm is shown in Fig. 2 (c). The key di\u000berence be-\ntween Fig. 2 (a) and (c) is the feature in (c) that extends\nup from the lower left-hand corner. This is highlighted in\nFig. 2 (d) where the data from (c) is overlaid with a solid\nblue curve showing the YIG ferromagnetic resonance con-\ndition. The blue dots show the YIG resonance condition\nas measured by re\rected microwave power. These data\nare \ft (blue curve) using the equation for the uniform\nFMR mode in a thin \flm [27], and we obtain a magneti-\nzation,\u00160Ms, of 183 mT (see supplementary information\nfor more information). As can be seen, the intensity of\nthe PL from the NV centers strongly changes precisely\nwhen the YIG FMR is excited.\nWe have considered the e\u000bect of heating, caused by the\nFMR absorption in YIG, on the PL of NV centers. Sev-\neral control experiments and estimates of the possible\ne\u000bect render this potential explanation highly unlikely.\nMore details can be found in the supplementary infor-\nmation.\nThere are two key points to note about the FMR-\ninduced feature in the PL signal. First, it is seen at fre-\nquencies and \felds well separated from the NV center's\nown resonance conditions. This is in clear contrast to\nthe quantum computing and magnetometry techniques\nbeing developed, where the spin-state of NV centers is\ncoherently manipulated by microwave \felds meeting the\nmagnetic resonance conditions [9{20]. Instead, we see a\nchange in the PL that correlates to the excitation of the\n3020100\nMagnetic Field (mT)\n3020100\nMagnetic Field (mT)4.0\n3.5\n3.0\n2.5\n2.0\n1.5\n1.0\n0.5Microwave Frequency (GHz)\n4.0\n3.5\n3.0\n2.5\n2.0\n1.5\n1.0\n0.5Microwave Frequency (GHz)\n(a) (b)\n(c) (d) NV || H \n NV /s94 H \n YIG data\n Fit\nOn YIGOn GGG\n1.00.50.0\nChange in NV PL (a.u.)FIG. 2. NV PL Data while YIG undergoes FMR : (a)\nRaw data showing the change in the intensity of the PL from\nthe NV centers in nanodiamonds dispersed on top of GGG,\nwith no YIG, as a function of microwave frequency and mag-\nnetic \feld. (b) The same data overlaid with theoretical reso-\nnance conditions for the NV centers. The black dashed lines\nshow the resonance condition for an NV center with the mag-\nnetic \feld parallel to the NV axis for the ground and excited\nstates. The grey crosses show the resonance condition for an\nNV center with the magnetic \feld perpendicular to the NV\naxis. (c) Similar data from nanodiamonds dispersed on the\nYIG sample with the distinct feature corresponding to the\nYIG FMR condition. (d) The data from (c) with the FMR\npeaks measured using re\rected microwave power (blue dots).\nAlso shown is a \ft (blue line) to the calculated dispersion re-\nlation for YIG \flm FMR with the magnetic \feld in plane (see\nmain text).3\nFIG. 3. NV center PL Data on YIG and on Wire\nAbove YIG : (a) Data taken with the laser spot focused on\ntop of the \u0018300 nm thick patterned microwire. The data\nshows that the signal corresponding to the YIG FMR is still\npresent, though reduced. The coupling must extend at least\n300 nm through the wire. (b) Data from the nanodiamonds\ndirectly on top of the YIG for comparison. (c) Line cuts of\nthe data in (a) and (b) at 16.5 mT. The FMR-induced peak is\nreduced by 3 times relative to the NV excited state peak for\nthe case of nanodiamonds on top of the microstrip, compared\nto the the case directly on YIG.\nYIG magnetization into precession by means of ferro-\nmagnetic resonance. It is remarkable that excitations at\nenergies as much as three to six times smaller than any\nNV center resonance have such a large e\u000bect on the NV\ncenter spin state. Second, the FMR-induced feature is\ncomparable in amplitude to the intrinsic NV resonances.\nThe large amplitude of the signal implies that a signif-\nicant number of NV centers in our laser spot must be\ncontributing to the signal. This suggests that since the\nnanodiamond \flm is 500 nm thick, the coupling must\nbe either long range (extending hundreds of nanometers)\nor that spin transport by means of spin di\u000busion plays a\nrole [28].\nTo probe the spatial extent of the coupling, we re-\npeated the measurement with our laser spot focused on\nthe nanodiamonds on top of the microwire, where the\nnanodiamonds are separated from the YIG by more than300 nm. This data can be seen in Fig. 3 (a) and compared\nto the signal when the diamond is directly on top of the\nYIG in Fig. 3 (b). Linecuts (at 16.5 mT), presented in\nFig. 3 (c), show that the FMR-induced feature is reduced\nbut clearly persists even when the nanodiamonds are not\nin direct contact with the YIG.\nWhile a clear explanation for the e\u000bect is not forthcom-\ning from our experimental results, we can make a few\nobservations. First, the insulating nature of both ma-\nterials rules out long range carrier-mediated transport\nof spins. Second, one of the unique aspects of this ex-\nperiment is that both the YIG magnetization and the\nNV center spins are out of equilibrium when the YIG\nis on FMR: YIG due to its resonance and the NV cen-\nters due to the hyperpolarizing action of the laser ex-\ncitation [29, 30]. This is in contrast to other systems\nwhere angular momentum is transferred between spin\nsub-systems, e.g. dynamic nuclear polarization or spin\npumping [1, 2, 31], where spins in one spin sub-system\nrelax by transferring polarization to another spin sub-\nsystem. The transfer of angular momentum from the\nYIG to the NV centers could result in relaxation towards\nequilibrium of both the spin systems, and thus be highly\ndesirable for the overall system.\nThis phenomenon o\u000bers the opportunity of probing\nspin transport in the absence of conductors and employ-\ning an all-optical readout. These advantages dramat-\nically reduce the potential confounding e\u000bects encoun-\ntered in studies of spin transport in metallic systems.\nThus this system o\u000bers an attractive and powerful ap-\nproach to better understanding microscopic details of re-\nlaxation and spin transport in magnetic heterostructures.\nExperiments are underway on other ferromagnets and\nstructures to gain more insight on this e\u000bect and how\nangular momentum is transferred between spin systems.\nIn summary, we have shown that the NV center spin\nstate can be manipulated by a dynamic coupling to the\nmagnetization of YIG. The availability of a wide selec-\ntion of ferromagnetic materials and structures (e.g., satu-\nration magnetization and anisotropies) potentially o\u000bers\na high degree of control in manipulating the NV spin\nstate. The potential for ultra high resolution imaging\nof ferromagnetic phenomena using individual NV centers\ncan have a signi\fcant impact as well. It should in\ru-\nence the \felds of spintronics and quantum information\nby combining the sensitivity of the NV center and the\ntunability and scope of ferromagnetism.\nThe authors wish to thank Yaroslav Tserkovnyak for\nuseful discussions. Funding for this research was provided\nby the Center for Emergent Materials at the Ohio State\nUniversity, an NSF MRSEC (Award Number DMR-\n0820414, ARO (Award number W911NF-12-1-0587) and\nDOE (Award number DE-FG02-03ER46054).4\n\u0003hammel@physics.osu.edu\n[1] C. P. Slichter, Principles of Magnetic Resonance\n(Springer-Verlag, New York, 1989).\n[2] F. Meier and B. P. Zakharchenya, Optical Orientation\n(North-Holland, Amsterdam, 1984).\n[3] P. C. Hammel, M. L. Roukes, Y. Hu, T. J. Gramila,\nT. Mamiya, and R. C. Richardson, Phys. Rev. Lett. 51,\n2124 (1983).\n[4] F. Pulizzi, ed., Nat. Mater. (Insight Issue:Spintronics) ,\nVol. 11 (Nature Publishing Group, 2012).\n[5] I. \u0014Zuti\u0013 c, J. Fabian, and S. D. Sarma, Rev. Mod. Phys.\n(2004).\n[6] D. D. Awschalom and M. E. Flatte, Nat. Phys. 3, 153\n(2007).\n[7] E. van Oort, N. Manson, and M. Glasbeek, Journal of\nPhysics C: Solid State Physics 21, 4385 (1988).\n[8] A. Gruber, A. Drabenstedt, C. Tietz, L. Fleury,\nJ. Wrachtrup, and C. vonBorczyskowski, Science 276,\n2012 (1997).\n[9] G. Balasubramanian, I. Y. Chan, R. Kolesov, M. Al-\nHmoud, J. Tisler, C. Shin, C. Kim, A. Wojcik, P. R. Hem-\nmer, A. Krueger, T. Hanke, A. Leitenstorfer, R. Brats-\nchitsch, F. Jelezko, and J. Wrachtrup, Nature 455, 648\n(2008).\n[10] M. V. Dutt, L. Childress, L. Jiang, E. Togan, J. Maze,\nF. Jelezko, A. S. Zibrov, P. R. Hemmer, and M. D. Lukin,\nScience 316, 1312 (2007).\n[11] G. D. Fuchs, G. Burkard, P. V. Klimov, and D. D.\nAwschalom, Nature Physics 7, 789 (2011).\n[12] M. S. Grinolds, S. Hong, P. Maletinsky, L. Luan, M. D.\nLukin, R. L. Walsworth, and A. Yacoby, Nature Physics\n9, 215 (2013).\n[13] S. Kaufmann, D. A. Simpson, L. T. Hall, V. Perunicic,\nP. Senn, S. Steinert, L. P. McGuinness, B. C. Johnson,\nT. Ohshima, F. Caruso, J. Wrachtrup, R. E. Scholten,\nP. Mulvaney, and L. Hollenberg, Proceedings of the Na-\ntional Academy of Sciences 110, 10894 (2013).\n[14] D. Le Sage, K. Arai, D. R. Glenn, S. J. DeVience, L. M.\nPham, L. Rahn-Lee, M. D. Lukin, A. Yacoby, A. Komeili,\nand R. L. Walsworth, Nature 496, 486 (2013).\n[15] P. Maletinsky, S. Hong, M. S. Grinolds, B. Hausmann,\nM. D. Lukin, R. L. Walsworth, M. Loncar, and A. Ya-\ncoby, Nat Nanotechnol 7, 320 (2012).\n[16] H. J. Mamin, M. Kim, M. H. Sherwood, C. T. Rettner,K. Ohno, D. D. Awschalom, and D. Rugar, Science 339,\n557 (2013).\n[17] J. R. Maze, P. L. Stanwix, J. S. Hodges, S. Hong, J. M.\nTaylor, P. Cappellaro, L. Jiang, M. V. G. Dutt, E. Togan,\nA. S. Zibrov, A. Yacoby, R. L. Walsworth, and M. D.\nLukin, Nature 455, 644 (2008).\n[18] S. Steinert, F. Ziem, L. T. Hall, A. Zappe, M. Schweikert,\nN. Gotz, A. Aird, G. Balasubramanian, L. Hollenberg,\nand J. Wrachtrup, Nature Communications 4(2013),\nArtn 1607 Doi 10.1038/Ncomms2588.\n[19] J. M. Taylor, P. Cappellaro, L. Childress, L. Jiang,\nD. Budker, P. R. Hemmer, A. Yacoby, R. Walsworth,\nand M. D. Lukin, Nature Physics 4, 810 (2008).\n[20] N. Bar-Gill, L. M. Pham, A. Jarmola, D. Budker, and\nR. L. Walsworth, Nature Communications 4(2013), Artn\n1743 Doi 10.1038/Ncomms2771.\n[21] D. M. Toyli, C. F. de las Casas, D. J. Christle, V. V.\nDobrovitski, and D. D. Awschalom, Proceedings of the\nNational Academy of Sciences of the United States of\nAmerica 110, 8417 (2013).\n[22] V. Cherepanov, I. Kolokolov, and V. Lvov, Physics\nReports-Review Section of Physics Letters 229, 81\n(1993).\n[23] A. A. Serga, A. V. Chumak, and B. Hillebrands, Journal\nof Physics D: Applied Physics 43, 264002 (2010).\n[24] H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel,\nand F. Y. Yang, Phys. Rev. B 88, 100406 (2013).\n[25] C. H. Du, H. L. Wang, Y. Pu, T. Meyer, P. Woodward,\nF. Y. Yang, and P. C. Hammel, Phys. Rev. Lett. 111,\n247202 (2013).\n[26] H. L. Wang, C. H. Du, Y. Pu, R. Adur, P. C. Hammel,\nand F. Y. Yang, arXiv:1307.2648v2 (2013).\n[27] C. Kittel, \\Introduction to solid state physics,\" (John\nWiley and Sons, Inc., Hoboken, NJ, 2005) Chap. Mag-\nnetic Resonance, eighth ed.\n[28] J. Cardellino, N. Scozzaro, M. R. Herman, A. J. Berger,\nC. Zhang, K. C. Fong, C. Jayaprakash, D. V. Pelekhov,\nand P. C. Hammel, \\Observation of Pure Spin Trans-\nport in a Diamond Spin Wire,\" (2014), arXiv:1309.3199\n[cond-mat.mes-hall].\n[29] J. H. N. Loubser and J. A. v. Wyk, Reports on Progress\nin Physics 41, 1201 (1978).\n[30] J. Harrison, M. Sellars, and N. Manson, Journal of Lu-\nminescence 107, 245 (2004).\n[31] Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys.\nRev. Lett. 88, 117601 (2002)." }, { "title": "1506.05879v2.Coherent_manipulation_of_a_Majorana_qubit_by_a_mechanical_resonator.pdf", "content": "Coherent manipulation of a Majorana qubit by a mechanical resonator\nP. Zhang1and Franco Nori1, 2\n1CEMS, RIKEN, Saitama 351-0198, Japan\n2Department of Physics, The University of Michigan, Ann Arbor, Michigan 48109-1040, USA\n(Dated: March 21, 2022)\nWe propose a hybrid system composed of a Majorana qubit and a nanomechanical resonator,\nimplemented by a spin-orbit-coupled superconducting nanowire, using a set of static and oscillating\nferromagnetic gates. The ferromagnetic gates induce Majorana bound states in the nanowire, which\nhybridize and constitute a Majorana qubit. Due to the oscillation of one of these gates, the Majorana\nqubit can be coherently rotated. By tuning the gate voltage to modulate the local spin-orbit\ncoupling, it is possible to reach the resonance of the qubit-oscillator system for relatively strong\ncouplings.\nPACS numbers: 71.10.Pm, 07.10.Cm\nI. INTRODUCTION\nMajorana bound states (MBS)1,2are recently attract-\ning increasing interest both theoretically and experimen-\ntally. These have been predicted to exist in arti\fcial\nstructures, such as nanowires with spin-orbit coupling\n(SOC) in proximity to a superconductor,3,4ferromag-\nnetic atom chains on top of a superconductor,5topo-\nlogical insulator/superconductor hybrid structures,6{11\nand superconducting circuits.12Recently, possible sig-\nnatures of MBS have been reported in nanowires,13{15\natom chains,16and topological insulator/superconductor\nstructures.17The MBS attract considerable atten-\ntion partly due to their hypothetical non-abelian\nanyonic statistics, which might allow the realiza-\ntion of topologically-protected quantum information\nmanipulation.2,18{20In parallel to the ongoing search of\nsome unambiguous con\frmation of MBS,21,22there are\nalso numerous theoretical studies on how to e\u000eciently\nexploit these MBS.\nOne promising application of MBS is to construct Ma-\njorana qubits.18It has been suggested that Majorana\nqubits might be robust against local perturbations and\nare hence promising to store quantum information.18,23,24\n(Note that Majorana qubits are not totally protected\nfrom decoherence, as studied in, e.g., Refs. 25{28.)\nFurthermore, Majorana qubits could be rotated by\ntopologically-protected braiding operations.19,29There-\nfore, among various realizations of qubits,30{36Majo-\nrana qubits are considered to be promising candidates for\nbuilding blocks of quantum information processors. The\nbraiding operations alone are insu\u000ecient to realize a uni-\nversal quantum gate based on a Majorana qubit.18For\nthe implementation of arbitrary qubit rotations, other\nnon-topological operations are required. Several schemes\nof such non-topological operations assisted by, e.g., phase\ngates,37,38quantum dots,39,40\rux qubits,41,42or mi-\ncrowave cavities,43have been proposed in the literature.\nNanomechanical resonators44could also be used to\nstudy non-topological operations of a Majorana qubit.\nFor example, quite recently, Kovalev et al.45have pro-posed to rotate a Majorana qubit by a magnetic can-\ntilever. Indeed, nanomechanical resonators have been\nutilized to couple to a wide range of quantum systems,\nincluding electric circuits,46optomechanical devices,47\natoms,48Cooper-pair boxes,49spin qubits,50or mi-\ncrowave cavities.51With the assistance of nanomechani-\ncal resonators, it is possible to perform important appli-\ncations such as quantum manipulations, quantum mea-\nsurements, as well as e\u000ecient sensing. These applications\nexploit the advantages of nanomechanical resonators,\ne.g., their large quality factors (103-106), high natural\nfrequencies (MHz-GHz), as well as the feasibility of reach-\ning the quantum ground states by cooling methods.52{54\nRecently, nanomechanical resonators have also been ex-\nploited to measure or manipulate the MBS.55{57Never-\ntheless, the study of hybrid systems58coupling nanome-\nchanical resonators to Majorana qubits is quite limited.\nThis work aims to contribute to this \feld. In this paper,\nwe propose another Majorana qubit-nanomechanical res-\nonator hybrid system in the framework of the spin-boson\nmodel,59based on a semiconductor nanowire in proxim-\nity to an s-wave superconductor. We show that a strong\ncoupling between a nanomechanical resonator and a Ma-\njorana qubit can be achieved, allowing an e\u000ecient trans-\nfer of quantum information between these two quantum\nsystems. Further, with braiding operations, it should be\npossible to realize a universal quantum gate based on a\nMajorana qubit.\nThis paper is organized as follows. First, we describe\nthe Majorana qubit and its coupling to a nanomechani-\ncal resonator. Afterwards, we numerically study the cou-\npling strength and the resonance condition of the hybrid\nsystem. Then, we solve the qubit-phonon dynamics and\nachieve a coherent control of the Majorana qubit. Fi-\nnally, we summarize our results.\nII. MODEL AND HAMILTONIAN\nAs illustrated in Fig. 1, we consider a semiconductor\nnanowire with a Rashba SOC of strength \u000bR\n0on the sur-\nface of an s-wave superconductor with a superconductingarXiv:1506.05879v2 [cond-mat.mes-hall] 24 Aug 20152\ngap \u0001. Three ferromagnetic gates, FM1, FM2 and FM3,\nare placed on top of and along the nanowire. Among\nthese gates, FM1 and FM3 are static while FM2 is free\nto harmonically oscillate along the nanowire (with a mass\nMand an oscillation frequency !0). The gates FM1\nand FM3 are su\u000eciently long (of the order of 1-10 \u0016m)\nwhile FM2 in between is relatively short (of the order of\n100 nm). These ferromagnetic gates induce a local Zee-\nman splitting in the nanowire. For simplicity, we take\nthe Zeeman splitting under the three gates to be identi-\ncal, with a magnitude of B0. An electric voltage Vcan\nbe applied on the gates to modulate the Rashba SOC lo-\ncally, e.g., from \u000bR\n0to\u000bR\nV. In our study, we consider the\ncase withB2\n0>(\u00012+\u00162), where\u0016is the chemical poten-\ntial in the nanowire. Therefore in the nanowire, the parts\nsubject to the Zeeman splitting (under the three ferro-\nmagnetic gates) are in the topological ( T) region. The\nremaining parts, without the Zeeman splitting, are in the\nnon-topological ( N) region.4,19As a result, the nanowire\nhas anN-T-N-T-N-T-Ndomain structure where the\nthreeTdomains are under the gates. At the six bound-\naries between the NandTdomains in the nanowire,\nMBS arise. As the two outer MBS are far apart, only the\nfour inner ones, schematically labeled as \r1-\r4in Fig. 1,\nare coupled due to hybridization arising from their small\nseparation60{63and are hence relevant to our considera-\ntion.\nTo lowest order, the hybrid system constructed above\ncan be described by the Hamiltonian\nH=HM+Hosc; (1)\nwhere the mutual coupling Hamiltonian of the MBS60{63\nHM=ign[l12(t)]\r1\r2+igt(l23)\r2\r3+ign[l34(t)]\r3\r4;\n(2)\nand the nanomechanical oscillator Hamiltonian\nHosc=p2\n2M+1\n2M!2\n0x2\n0(t): (3)\nThe coupling strengths gn;tdepend on the domain\nlengthslij. Due to the oscillation of the gate FM2,\nl12(t) =l0\n12+x0(t) andl34(t) =l0\n34\u0000x0(t) are time\ndependent. Here, x0(t) stands for the displacement of\nthe gate FM2 from its balance position, which is much\nsmaller than the static domain lengths l0\n12;34. Therefore,\nto \frst order in x0, one has\nHM=i[gn(l0\n12) +x0(t)g0\nn(l0\n12)]\r1\r2+igt(l23)\r2\r3\n+i[gn(l0\n34)\u0000x0(t)g0\nn(l0\n34)]\r3\r4: (4)\nThe four MBS, satisfying f\ri;\rjg=\u000eij, can be used\nto construct a Majorana qubit as follows.2,18At \frst we\nde\fne two Dirac fermion operators64c\"= (\r1+i\r4)=p\n2\nandc#= (\r2+i\r3)=p\n2. The Hilbert space of HMcan\nthen be spanned by states jn\";n#i, with the fermion oc-\ncupation numbers n\"=cy\n\"c\"andn#=cy\n#c#. Due to the\nxy\nγ1γ2γ3γ4\nFM1\nFM3\nsuperconductorl\n 12\nFM2\nω0\nV(a)\nl\n 23l\n 34\n(b)\nx(µm)\nB/B 0|Ψ|2(arb. units)2\n1\n0\n1 0 -13\n2\n1\n0FIG. 1: (Color online) (a) Schematic diagram of the proposed\nMajorana qubit-nanomechanical resonator hybrid system. A\nsemiconductor nanowire is placed on the surface of an s-wave\nsuperconductor. Three ferromagnetic gates are on top of the\nnanowire, of which FM1 and FM3 are su\u000eciently long (of the\norder of 1-10 \u0016m) and static, while FM2 is relatively short (of\nthe order of 100 nm) and free to oscillate as a harmonic oscil-\nlator. The ferromagnetic gates induce a local Zeeman split-\ntingB0in the underlying nanowire, and can also be used to\nmodulate the local Rashba SOC strength by applying an elec-\ntric voltage V. (b) Wave amplitude j\tj2of the four coupled\nMBS at a static state in an InSb nanowire with the set-up\nshown in (a). The red dashed and red solid curves respec-\ntively correspond to the wave amplitudes of the lowest two\neigenstates (close to the zero energy). These two states con-\nstitute the Majorana qubit. The dotted curve with the scale\non the right-hand side of the frame indicates the pro\fle of the\ninhomogeneous Zeeman splitting along the nanowire. In the\ncalculation l12=l34= 150 nm, l23= 400 nm, B0= 1 meV,\nand \u0001 = 0:5 meV. The gate voltage Vis zero and the Rashba\nSOC is homogeneous along the nanowire, with a strength\n\u000bR\n0= 20 meV nm.\nconservation of fermion parity, the states fj0;1i;j1;0ig\nandfj0;0i;j1;1igform two decoupled (odd and even)\nsectors.2,41,43We assume that there is no high-energy\nexcitation (e.g., no Cooper-pair breaking in the super-\nconducting substrate) and restrict our study to the odd\nsector with n\"+n#= 1. For convenience, we de\fne\npseudo-spinsj\"i=j1;0iandj#i=j0;1i, and use them\nas the two logical states of the Majorana qubit.39,41,43,45\nIn this pseudo-spin space, i\r1\r2=\u0000i\r3\r4=\u0000\u001byand\ni\r2\r3=\u0000\u001bz. The nanomechanical oscillator is quan-3\ntized in the Fock space fjnigwithjni=(ay)n\np\nn!j0i, where\na=q\nM! 0\n2~(x0+i\nM! 0p) is the annihilation operator of\nphonons. Consequently, in the space fj\"#i\njnig, the\nhybrid system can be simply described by the spin-boson\nHamiltonian,59\nHe\u000b=\u0000\"\n2\u001bz\u0000\u000e\u001by+g(ay+a)\u001by+n~!0; (5)\nwith the constant omitted. Here \"= 2gt(l23),\u000e=\ngn(l0\n12)\u0000gn(l0\n34), andg=\u0000~x0[g0\nn(l0\n12) +g0\nn(l0\n34)], where\n~x0= [~=(2M! 0)]1=2is the zero-point motion of the oscil-\nlator.\nIII. HYBRIDIZATION OF MAJORANA BOUND\nSTATES\nIn this section, we study the MBS and their mu-\ntual coupling. In the static state, the inhomoge-\nneous nanowire can be described by a tight-binding\nmodel. Using the Bogoliubov-de Gennes basis \t j=\n(fj\";fj#;fy\nj#;\u0000fy\nj\"), wherefj\u0011stands for the fermion op-\nerator of a spin- \u0011(\u0011=\",#) electron on the j-th lattice\nsite, the particle-hole Hamiltonian reads5\nHBDG =1\n2X\nj[\ty\nj^hj\tj+ (\ty\nj^tj\tj+1+ H:c:)];(6)\nwhere\n^hj= (2t0\u0000\u0016)s0\u001cz+ \u0001s0\u001cx+Bjsy\u001c0; (7)\n^tj=t0s0\u001cz+i\u000bjsz\u001cz: (8)\nIn the above Hamiltonian, the Pauli matrices \u001cx;y;z act\non the particle-hole space and sx;y;z act on the real\nspin space. The spin-diagonal hopping energy is t0=\n~2=(2m\u0003a2), and the spin-o\u000b-diagonal hopping energy is\n\u000bj=\u000bR\nj=(2a). HereBjand\u000bR\njare the on-site Zeeman\nsplitting and Rashba SOC, respectively, m\u0003is the e\u000bec-\ntive electron mass, and ais the lattice spacing in the\ndiscretized tight-binding model. In the T(N) domains\nBj=B0(Bj= 0) and\u000bR\nj=\u000bR\nV(\u000bR\nj=\u000bR\n0). When the\ngate voltage Vis zero,\u000bR\nV=\u000bR\n0.\nHere, to lowest order, we follow Refs. 45 and 11 to\ninvestigate the coupling strength gn(gt) approximately\nin an isolated T-N-T(N-T-N) three-domain structure.\nIn such a simpli\fed model, the inner N(T) domain has\na \fnite length, while the outer two T(N) domains are\nassumed to be in\fnitely long. By numerically diagonal-\nizing this three-domain system, one can obtain the en-\nergy splitting of the two MBS localized at the two T/N\nboundaries. This energy splitting is precisely caused by\nthe coupling of the MBS. With gnandgtknown numeri-\ncally, the Majorana qubit can be well described by \"and\n\u000e, and the qubit-phonon coupling gcan be obtained also\nfromg0\nn[refer to Eq. (5)]. Moreover, by exactly diagonal-\nizing the Hamiltonian of the genuine N-T-N-T-N-T-Ndomain structure as shown in Fig. 1(a), one can obtain\nthe hybrid four MBS under consideration.\nIn this work, we consider an InSb quantum wire13with\nan e\u000bective electron mass m\u0003= 0:015me, a Rashba SOC\n\u000bR\n0= 20 meV nm, and a large Landau factor gL\u001950.\nWe choose the superconducting gap \u0001 = 0 :5 meV, the\nlocal Zeeman splitting B0= 1 meV, the chemical poten-\ntial\u0016= 0, and the lattice constant a= 10 nm. The total\nlattice site number is chosen as 1000 for the numerical\nconvergence. In Fig. 2, we show the dependence of the\nMajorana coupling strength gn(gt) on the length of the\nN(T) domainln(lt), as well as the derivative g0\nnver-\nsusln. Further, as an example, in Fig. 1(b) we present\nthe wave amplitude j\tj2of the four hybrid MBS, when\nl12=l34= 150 nm and l23= 400 nm. In Fig. 1(b), the\nred dashed and red solid curves stand for the wave am-\nplitudes of the lowest two eigenstates (close to the zero\nenergy) in the static inhomogeneous nanowire. The state\ncorresponding to the red solid (dashed) curve is mainly\ncontributed by the \r2and\r3(\r1and\r4) MBS. Here,\nthese two states form the Majorana qubit.\ng′\nngtgn\nln(t)(nm)\ng′\nn(µeV/nm)gn(t)(µeV)0\n-0.5\n-1\n-1.5\n-2\n500 400 300 200 100 0100\n0\n-100\n-200\nFIG. 2: (Color online) Majorana coupling strength gn(gt)\nversusln(lt), the length of the inner N(T) domain between\nthe two outer T(N) domains. The derivative g0\nnversuslnis\nalso shown, with the scale on the right hand side of the frame.\nThe necessary parameters for the calculation are speci\fed in\nthe main text.\nIV. QUBIT-PHONON COUPLING AND\nRESONANCE\nWe now look into the qubit-phonon coupling and the\nresonance condition. We assume that the nanomechan-\nical oscillator FM2 has a mass M= 10\u000015Kg and an\noscillation frequency !0= 5 MHz. With these parame-\nters, the zero-point motion of the oscillator is calculated\nto be ~x0= 0:1 pm. We consider the symmetric case\nwithl0\n12=l0\n34= 150 nm, and hence we have \u000e= 0 and\ng= 0:2 MHz in Eq. (5). The longitudinal length l23of\nthe FM2 gate is chosen as 400 nm, such that the Rabi4\nresonance condition \"\u0019\u0000!0can be easily satis\fed, e.g.,\nby further subtly adjusting the gate voltage Vwhich con-\ntrols the local Rashba SOC strength \u000bR\nV. In Fig. 3, we\npresent the variation of \"as well asgversus\u000bR\nV. It is\nshown that when slightly adjusting V, and hence \u000bR\n0,\nthe resonance point \"\u0019\u0000!0can be reached while the\nqubit-phonon coupling gremains almost invariant. This\nqubit-phonon coupling is relatively strong, in view of the\nlong lifetime of the Majorana qubit and the high qual-\nity factor of the nanomechanical oscillator. In principle,\nthe qubit-phonon coupling can be stronger when the do-\nmain length l12(as well as l34) becomes smaller (refer\nto Fig. 2). However, if the two edge modes \r1and\r2\n(as well as \r3and\r4) are too close and hence their hy-\nbridization becomes quite strong, the model Hamiltonian\n(2) describing four distinguishable MBS might fail.\ng/ω 0ε/ω 0\nresonance\nαR\nV/αR\n0\ng/ω 0ε/ω 00.1\n0\n-0.1\n1.0685 1.068 1.067520\n10\n0\n-10\n-20\nFIG. 3: (Color online) Qubit splitting \"and the qubit-phonon\ncouplingg(the scale is on the right-hand side of the frame)\nversus\u000bR\nV, the Rashba SOC strength in the topological ( T)\ndomains modulated by the gate voltage V.\nV. QUBIT-PHONON DYNAMICS\nHere we study the dynamics of the qubit-phonon hy-\nbrid system. To achieve this, we make use of the Python-\nbased Qutip software package65,66to solve the Lindblad\nmaster equation,\n_\u001a(t) =\u0000i\n~[He\u000b;\u001a(t)] +1\n2X\nkn\n[Lk;\u001a(t)Ly\nk]\n+ [Lk\u001a(t);Ly\nk]o\n: (9)\nIn this equation, \u001ais the density matrix of the qubit-\nphonon system, and Lkare the Lindblad operators ac-\ncounting for the dissipation of the hybrid system due to\nits coupling to the environment. The relaxation of the\nMajorana qubit is taken into account by L1=p\n1=T1\u001b\u0000,\nwhile the dissipation of the nanomechanical resonator isincluded by L2=p\n(\u0016n+ 1)!0=Qa andL3=p\n\u0016n!0=Qay.\nHere \u0016n= [exp( ~!0=kB~T)\u00001]\u00001is the thermal phonon\nnumber in equilibrium with the environmental tempera-\nture ~T,Qis the quality factor of the nanomechanical os-\ncillator, and T1is the usual relaxation time of the qubit.\nBy solving the master equation, one can obtain the time\nevolution of the qubit and phonon occupations.\nIn our model, the temperature ~Tis set as 10 mK and\nhence the thermal phonon number \u0016 nis as large as 258.\nTherefore, an additional cooling of the oscillator52{54is\nrequired, e.g., as also applied in a proposed nanomechan-\nical resonator{nitrogen-vacancy center hybrid system.51\nWe assume that after side-band cooling52{54the phonons\nthermally occupy the lowest several quantum states with\na small phonon number, e.g., n= 0:3. The initial state of\nthe Majorana qubit is set as j\"i, implying that a single\nelectron is splitted into the \r1and\r4Majorana fermions.\nExperimentally, this initial state might be realized when\nonly the FM1 and FM3 gates are in proximity to the\nnanowire before inserting the middle FM2 gate. The re-\nlaxation time of the Majorana qubit depends on the con-\ncrete set-up and environment. Following Refs. 26 and 27,\nwe typically set T1around 100 \u0016s.\nIn Fig. 4, we plot the time evolution of the occupations\nof the qubit and phonons respectively, with di\u000berent val-\nues ofT1andQ. As indicated by the \fgure, quantum\ninformation can be e\u000bectively transferred back and forth\nbetween the Majorana qubit and the nanomechanical res-\nonator. During this process, the single electron in the\nnanowire alternatively occupies (back and forth) the pair\nof MBS:\r1and\r4, or\r2and\r3. Inversely, this quan-\ntum information transfer can also modulate the motion of\nthe oscillator, e.g., the oscillation amplitude. In fact, as\nthe nanomechanical resonator is near its quantum ground\nstate, the oscillation amplitude hx2\n0i, which might be ob-\nservable, is almost linearly related to the phonon number.\nThis is becausehx2\n0i/h (ay+a)2i\u00192hayai+ 1 = 2n+ 1.\nTherefore, the dashed curves in Fig. 4, representing the\ntime evolution of the phonon number, also supply in-\nformation on the change of the oscillation amplitude of\nthe resonator due to its coupling with the qubit. This\nphenomenon signi\fes the presence of a Majorana qubit.\nCertainly, for better performance of this hybrid system\n(e.g., with a higher \fdelity), a higher quality factor of\nthe resonator and a longer relaxation time of the qubit\nare preferred.\nVI. DISCUSSION\nHere we brie\ry compare our model to the one pro-\nposed by Kovalev et al. ,45where a vibrating cantilever is\nutilized to rotate a Majorana qubit. The e\u000bective Hamil-\ntonian in their model [Eq. (7) in Ref. 45] is in fact equiv-\nalent to the one in our manuscript [Eq. (5)]. This is\nunderstandable as both are in the framework of the spin-\nboson model. Note that for both cases there exists a\nstatic o\u000b-diagonal term [for our case, that is the \u000eterm5\n(b)T1= 150 µs\ngt(π)Occupation\n5 4 3 2 1 01\n0.5\n0(a)T1= 50 µs\nphononsQ= 3×105: qubitphononsQ= 106: qubit1.5\n1\n0.5\n0\nFIG. 4: (Color online) Time evolution of the occupations\nof the Majorana qubit (solid curves) and phonons (dashed\ncurves) which are in Rabi resonance. The qubit relaxation\ntime is set as 50 \u0016s in (a) and 150 \u0016s in (b). The calcula-\ntions for both (a) and (b) are performed with two di\u000berent\nresonator quality factors: Q= 106andQ= 3\u0002105.\nin Eq. (5)] coupling the two levels of the qubit in the\nHamiltonian. To neglect this term, in order to simplify\nthe theoretical analysis, some conditions have to be satis-\n\fed. Speci\fcally, in Ref. 5, a certain equilibrium angle ( \u00120\nthere) of the vibrating cantilever has to be established. In\nour opinion, exactly solving this angle and then adjust-\ning the experimental setup correspondingly45are chal-\nlenging. However, in order to neglect the constant o\u000b-\ndiagonal term in our case, the experimental setup must\nbe mirror-symmetric about the middle point of the FM2gate, i.e.,l0\n12=l0\n34. Therefore, we think that our model\nis more easily accessible by experiments and hence more\nadvantageous.\nVII. CONCLUSIONS\nIn conclusion, we have proposed a hybrid system com-\nposed of a Majorana qubit and a mechanical resonator,\nimplemented by a semiconductor nanowire in proxim-\nity to an s-wave superconductor. In this proposal, three\nferromagnetic gates are placed on top of and along the\nnanowire; the two outer gates are static and the inner\none is free to oscillate harmonically as a mechanical res-\nonator. These ferromagnetic gates induce a local Zeeman\nsplitting and give rise to four Majorana bound states,\nconstituting a Majorana qubit in the nanowire. The\ndynamical hybridization of the Majorana bound states,\narising from the motion of the oscillating gate, results in\na coherent coupling between the Majorana qubit and the\nmechanical resonator.\nThis hybrid system can be adjusted to be in resonance,\ne.g., with the assistance of a gate voltage on the ferro-\nmagnetic gates, which controls the Rashba SOC locally in\nthe nanowire. Our study reveals that under resonance, a\nstrong coupling between the qubit and the resonator can\nbe achieved. Consequently, quantum information can be\ne\u000bectively transferred from the Majorana qubit to the\noscillator and then back to the qubit. This quantum in-\nformation transfer can manifest itself in modulating the\nmotion of the oscillator, which may conversely signify the\npresence of the Majorana qubit.\nAcknowledgments\nThe authors gratefully acknowledge X. Hu, G. Gi-\navaras, L. Wang and Z. Li for valuable discussions\nand comments. P.Z. acknowledges the support of a\nJSPS Foreign Postdoctoral Fellowship under Grant No.\nP14330. F.N. is partially supported by the RIKEN\niTHES Project, MURI Center for Dynamic Magneto-\nOptics via the AFOSR award number FA9550-14-1-0040,\nthe IMPACT program of JST, and a Grant-in-Aid for\nScienti\fc Research (A).\n1A.Y. Kitaev, Unpaired Majorana Fermions in Quantum\nWires , Phys. Usp. 44, 131 (2001).\n2C.W.J. Beenakker, Search for Majorana Fermions in Su-\nperconductors , Annu. Rev. Con. Mat. Phys. 4, 113 (2013).\n3R.M. Lutchyn, J.D. Sau, and S. Das Sarma, Ma-\njorana Fermions and a Topological Phase Transition\nin Semiconductor-Superconductor Heterostructures , Phys.\nRev. Lett. 105, 077001 (2010).\n4Y. Oreg, G. Refael and F. von Oppen, Helical Liquids andMajorana Bound States in Quantum Wires , Phys. Rev.\nLett. 105, 177002 (2010).\n5T.P. Choy, J.M. Edge, A.R. Akhmerov, and C.W.J.\nBeenakker, Majorana Fermions Emerging from Magnetic\nNanoparticles on a Superconductor without Spin-Orbit\nCoupling , Phys. Rev. B 84, 195442 (2011).\n6L. Fu and C.L. Kane, Superconducting Proximity E\u000bect\nand Majorana Fermions at the Surface of a Topological\nInsulator , Phys. Rev. Lett. 100, 096407 (2008).6\n7A. Cook and M. Franz, Majorana Fermions in a\nTopological-Insulator Nanowire Proximity-Coupled to an\ns-Wave Superconductor , Phys. Rev. B 84, 201105 (2011).\n8A.L. Rakhmanov, A.V. Rozhkov, and F. Nori, Majorana\nFermions in Pinned Vortices , Phys. Rev. B 84, 075141\n(2011).\n9R.S. Akzyanov, A.V. Rozhkov, A.L. Rakhmanov, and F.\nNori, Tunneling Spectrum of a Pinned Vortex with a Ro-\nbust Majorana State , Phys. Rev. B 89, 085409 (2014).\n10R.S. Akzyanov, A.L. Rakhmanov, A.V. Rozhkov, and F.\nNori, Majorana Fermions at the Edge of Superconducting\nIslands , arXiv: 1504.05688.\n11L. Jiang, D. Pekker, J. Alicea, G. Refael, Y. Oreg, A.\nBrataas, and F. von Oppen, Magneto-Josephson E\u000bects\nin Junctions with Majorana Bound States , Phys. Rev. B\n87, 075438 (2013).\n12J.Q. You, Z.D. Wang, W. Zhang, and F. Nori, Encoding a\nQubit with Majorana Modes in Superconducting Circuits ,\nSci. Rep. 4, 5535 (2014).\n13V. Mourik, K. Zuo, S.M. Frolov, S.R. Plissard, and\nE.P.A.M. Bakkers, L. P. Kouwenhoven, Signatures of Ma-\njorana Fermions in Hybrid Superconductor Semiconductor\nNanowire Device , Science 336, 1003 (2012).\n14M.T. Deng, C.L. Yu, G.Y. Huang, M. Larsson, P. Caro\u000b,\nH.Q. Xu, Anomalous Zero-Bias Conductance Peak in a\nNb-InSb Nanowire-Nb Hybrid Device , Nano Lett. 12, 6414\n(2012).\n15A. Das, Y. Ronen, Y. Most, Y. Oreg, M. Heiblum, and\nH. Shtrikman, Zero-Bias Peaks and Splitting in an AlInAs\nNanowire Topological Superconductor as a Signature of\nMajorana Fermions , Nat. Phys. 8, 887 (2012).\n16S.N. Perge, I.K. Drozdov, J. Li, H. Chen, S. Jeon, J. Seo,\nA.H. MacDonald, B.A. Bernevig, and A. Yazdani, Ob-\nservation of Majorana Fermions in Ferromagnetic Atomic\nChains on a Superconductor , Science 346, 602 (2014).\n17J.P. Xu, M.X. Wang, Z.L. Liu, J.F. Ge, X.J. Yang, C.H.\nLiu, Z.A. Xu, D. Guan, C.L. Gao, D. Qian, Y. Liu, Q.H.\nWang, F.C. Zhang, Q.K. Xue, and J.F. Jia, Experimen-\ntal Detection of a Majorana Mode in the Core of a Mag-\nnetic Vortex inside a Topological Insulator-Superconductor\nBi2Te3/NbSe 2Heterostructure , Phys. Rev. Lett. 114,\n017001 (2015).\n18C. Nayak, S.H. Simon, A. Stern, M. Freedman, and S.\nDas Sarma, Non-Abelian Anyons and Topological Quan-\ntum Computation , Rev. Mod. Phys. 80, 1083 (2008).\n19J. Alicea, Y. Oreg, G. Refael, F. von Oppen, and M.P.A.\nFisher, Non-Abelian Statistics and Topological Quantum\nInformation Processing in 1D Wire Networks , Nat. Phys.\n7, 412 (2011).\n20S. Tewari, S. Das Sarma, C. Nayak, C. Zhang, and P.\nZoller, Quantum Computation Using Vortices and Majo-\nrana Zero Modes of a px+ipySuper\ruid of Fermionic Cold\nAtoms , Phys. Rev. Lett. 98, 010506 (2007).\n21J. Liu, A.C. Potter, K.T. Law, and P.A. Lee, Zero-\nBias Peaks in the Tunneling Conductance of Spin-Orbit-\nCoupled Superconducting Wires with and without Majo-\nrana End-States , Phys. Rev. Lett. 109, 267002 (2012).\n22E. Dumitrescu, B. Roberts, S. Tewari, J.D. Sau, and S.\nDas Sarma, Majorana Fermions in Chiral Topological Fer-\nromagnetic Nanowires , Phys. Rev. B 91, 094505 (2015).\n23L. Mao and C. Zhang, Robustness of Majorana Modes\nand Minigaps in a Spin-Orbit-Coupled Semiconductor-\nSuperconductor Heterostructure , Phys. Rev. B 82, 174506\n(2010).24J.D. Sau and S. Das Sarma, Realizing a Robust Practical\nMajorana Chain in a Quantum-Dot-Superconductor Lin-\near Array , Nat. Commun., 3, 964 (2012).\n25G. Goldstein and C. Chamon, Decay Rates for Topological\nMemories Encoded with Majorana Fermions , Phys. Rev. B\n84, 205109 (2011).\n26D. Rainis and D. Loss, Majorana Qubit Decoherence by\nQuasiparticle Poisoning , Phys. Rev. B 85, 174533 (2012).\n27M.J. Schmidt, D. Rainis, and D. Loss, Decoherence of Ma-\njorana Qubits by Noisy Gates , Phys. Rev. B 86, 085414\n(2012).\n28J.C. Budich, S. Walter, and B. Trauzettel, Failure of Pro-\ntection of Majorana based Qubits against Decoherence ,\nPhys. Rev. B 85, 121405 (2012).\n29D.A. Ivanov, Non-Abelian Statistics of Half-Quantum Vor-\ntices in p-Wave Superconductors , Phys. Rev. Lett. 86, 268\n(2001).\n30M.W. Wu, J.H. Jiang, and M.Q. Weng, Spin Dynamics in\nSemiconductors , Phys. Rep. 493, 61 (2010).\n31S.N. Perge, S.M. Frolov, E.P.A.M. Bakkers, and L.P.\nKouwenhoven, Spin-Orbit Qubit in a Semiconductor\nNanowire , Nature 468, 1084 (2010).\n32P. Zhang, Z.L. Xiang, and F. Nori, Spin-Orbit Qubit on\na Multiferroic Insulator in a Superconducting Resonator ,\nPhys. Rev. B 89, 115417 (2014).\n33R. Li, J.Q. You, C.P. Sun, and F. Nori, Controlling a\nNanowire Spin-Orbit Qubit via Electric-Dipole Spin Res-\nonance , Phys. Rev. Lett. 111, 086805 (2013).\n34I. Buluta and F. Nori, Quantum Simulators , Science, 326,\n108 (2009).\n35I. Buluta, S. Ashhab, and F. Nori, Natural and Arti\fcial\nAtoms for Quantum Computation , Rep. Prog. Phys. 74,\n104401 (2011).\n36J.Q. You and F. Nori, Atomic Physics and Quantum\nOptics Using Superconducting Circuits , Nature 474, 589\n(2011).\n37P. Bonderson, D.J. Clarke, C. Nayak, and K. Shtengel,\nImplementing Arbitrary Phase Gates with Ising Anyons ,\nPhys. Rev. Lett. 104, 180505 (2010).\n38D.J. Clarke and K. Shtengel, Improved Phase-Gate Relia-\nbility in Systems with Neutral Ising Anyons , Phys. Rev. B\n82, 180519 (2010).\n39K. Flensberg, Non-Abelian Operations on Majorana\nFermions via Single-Charge Control , Phys. Rev. Lett. 106,\n090503 (2011).\n40P. Bonderson and R.M. Lutchyn, Topological Quantum\nBuses: Coherent Quantum Information Transfer between\nTopological and Conventional Qubits , Phys. Rev. Lett.\n106, 130505 (2011).\n41D. Pekker, C.Y. Hou, V.E. Manucharyan, and E. Demler,\nProposal for Coherent Coupling of Majorana Zero Modes\nand Superconducting Qubits Using the 4 \u0019Josephson Ef-\nfect, Phys. Rev. Lett. 111, 107007 (2013).\n42F. Hassler, A.R. Akhmerov, C.Y. Hou, and C.W. J.\nBeenakker, Anyonic Interferometry without Anyons: How\na Flux Qubit Can Read Out a Topological Qubit , New J.\nPhys. 12, 125002 (2010).\n43T.L. Schmidt, A. Nunnenkamp, and C. Bruder, Majorana\nQubit Rotations in Microwave Cavities , Phys. Rev. Lett.\n110, 107006 (2013).\n44H.G. Craighead, Nanoelectromechanical Systems , Science\n290, 1532 (2000).\n45A.A. Kovalev, A. De, and K. Shtengel, Spin Transfer of\nQuantum Information between Majorana Modes and a7\nResonator , Phys. Rev. Lett. 112, 106402 (2014).\n46M. Blencowe, Quantum Electromechanical Systems , Phys.\nRep.395, 159 (2004).\n47M. Aspelmeyer, T.J. Kippenberg, and F. Marquardt, Cav-\nity Optomechanics , Rev. Mod. Phys. 86, 1391 (2014).\n48K. Hammerer, M. Wallquist, C. Genes, M. Ludwig, F. Mar-\nquardt, P. Treutlein, P. Zoller, J. Ye, and H.J. Kimble,\nStrong Coupling of a Mechanical Oscillator and a Single\nAtom , Phys. Rev. Lett. 103, 063005 (2009).\n49E.K. Irish and K. Schwab, Quantum Measurement of a\nCoupled Nanomechanical Resonator-Cooper-Pair Box Sys-\ntem, Phys. Rev. B 68, 155311 (2003).\n50P. Rabl, P. Cappellaro, M.V. Gurudev Dutt, L. Jiang,\nJ.R. Maze, and M.D. Lukin, Strong Magnetic Coupling\nbetween an Electronic Spin Qubit and a Mechanical Res-\nonator , Phys. Rev. B 79, 041302 (2009).\n51P.B. Li, Y.C. Liu, S.Y. Gao, Z.L. Xiang, P. Rabl, F.L.\nLi, and Y.F. Xiao, Hybrid Quantum Device based on NV\nCenters in Diamond Nanomechanical Resonators Plus Su-\nperconducting Waveguide Cavities , arXiv:1503.02437.\n52A.D. O'Connell, M. Hofheinz, M. Ansmann, R.C. Bialczak,\nM. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang,\nM. Weides, J. Wenner, J.M. Martinis, and A.N. Cleland,\nQuantum Ground State and Single-Phonon Control of a\nMechanical Resonator , Nature (London) 464, 697 (2010).\n53J.D. Teufel, T. Donner, D. Li, J.W. Harlow, M.S. All-\nman, K. Cicak, A.J. Sirois, J.D. Whittaker, K.W. Lehnert,\nand R.W. Simmonds, Sideband Cooling of Micromechan-\nical Motion to the Quantum Ground State , Nature 475,\n359 (2011).\n54J. Chan, T.P.M. Alegre, A.H.S. Naeini, J.T. Hill, A.\nKrause, S. Gr oblacher, M. Aspelmeyer, and O. Painter,\nLaser Cooling of a Nanomechanical Oscillator into Its\nQuantum Ground State , Nature 478, 89 (2011).\n55S. Walter, T.L. Schmidt, K. B\u001crkje, and B. Trauzettel, De-\ntecting Majorana Bound States by Nanomechanics , Phys.\nRev. B 84, 224510 (2011).\n56S. Walter and J.C. Budich, Teleportation-Induced Entan-\nglement of Two Nanomechanical Oscillators Coupled to\na Topological Superconductor , Phys. Rev. B 89, 155431(2014).\n57H.J. Chen and K.D. Zhu, All-Optical Scheme for Detect-\ning the Possible Majorana Signature based on QD and\nNanomechanical Resonator Systems , Science China 58,\n050301 (2015).\n58Z.L. Xiang, S. Ashhab, J.Q. You, and F. Nori, Hybrid\nQuantum Circuits: Superconducting Circuits Interacting\nwith Other Quantum Systems , Rev. Mod. Phys. 85, 623\n(2013).\n59A.J. Leggett, S. Chakravarty, A.T. Dorsey, M.P.A. Fisher,\nA. Garg, and W. Zwerger, Dynamics of the Dissipative\nTwo-State System , Rev. Mod. Phys. 59, 1 (1987).\n60C.J. Bolech and E. Demler, Observing Majorana Bound\nStates in p-Wave Superconductors Using Noise Measure-\nments in Tunneling Experiments , Phys. Rev. Lett. 98,\n237002 (2007).\n61G.W. Semeno\u000b and P. Sodano, Stretched Quantum States\nEmerging from a Majorana Medium , J. Phys. B 40, 1479\n(2007).\n62S. Tewari, C. Zhang, S. Das Sarma, C. Nayak, and D.-\nH. Lee, Testable Signatures of Quantum Nonlocality in\na Two-Dimensional Chiral p-Wave Superconductor , Phys.\nRev. Lett. 100, 027001 (2008).\n63Y.E. Kraus, A. Auerbach, H.A. Fertig, and S.H. Simon,\nTesting for Majorana Zero Modes in a px+ipySupercon-\nductor at High Temperature by Tunneling Spectroscopy ,\nPhys. Rev. Lett. 101, 267002 (2008).\n64Note that the pairing of MBS to form Dirac fermions is\narbitrary. A given pairing de\fnes a basis and di\u000berent bases\ncan be connected by a unitary transformation.18Here we\nde\fne Dirac fermions according to the initial state where a\nsingle electron is splitted into a pair of Majorana fermions\n\r1and\r4(refer to Sec. V).\n65J.R. Johansson, P.D. Nation, and F. Nori, QuTiP: A\nPython Framework for the Dynamics of Open Quantum\nSystems , Comput. Phys. Commun. 183, 1760 (2012).\n66J.R. Johansson, P.D. Nation, and F. Nori, QuTiP 2: A\nPython Framework for the Dynamics of Open Quantum\nSystems , Comput. Phys. Commun. 184, 1234 (2013)." }, { "title": "1809.06836v2.Local_manipulation_of_quantum_magnetism_in_1D_ultracold_Fermi_gases_across_narrow_resonances.pdf", "content": "arXiv:1809.06836v2 [cond-mat.quant-gas] 19 Sep 2018Local manipulation of quantum magnetism in 1D ultracold Fer mi gases across narrow\nresonances\nLei Pan,1,2Xiaoling Cui,1,∗and Shu Chen1,2,3, †\n1Beijing National Laboratory for Condensed Matter Physics,\nInstitute of Physics, Chinese Academy of Sciences, Beijing 100190, China\n2School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China\n3Collaborative Innovation Center of Quantum Matter, Beijin g, China\n(Dated: April 5, 2022)\nEffective range is a quantity to characterize the energy depe ndence in two-body scattering\nstrength, and is widely used in cold atomic systems especial ly across narrow resonances. Here we\nshow that the effective range can significantly modify the mag netic propertyof one-dimensional (1D)\nspin-1/2 fermions in the strongly repulsive regime. In particular, the effective range breaks the large\nspindegeneracy inthehard-corelimit, andinducesaHeisen bergexchangeterminthespinchainthat\nis much more sensitive to the local density than that induced by the bare coupling. With an external\nharmonic trap, this leads to a very rich magnetic pattern whe re the anti-ferromagnetic(AFM) and\nferromagnetic (FM) correlations can coexist and distribut e in highly tunable regions across the trap.\nFinally, we propose to detect the range-induced magnetic or der in the tunneling experiment. Our\nresults can be directly tested in 1D Fermi gases across narro w resonance, and suggest a convenient\nroute towards the local manipulation of quantum magnetism i n cold atoms.\nIntroduction. Energy dependent interaction is com-\nmon in nature, which roots deeply in the renormaliza-\ntion group theory. In cold atomic systems, such energy\ndependence appears naturally in the Feshbach resonance\nwhich essentially relies on the energy difference of (open)\natomic and (closed) molecular channels[ 1]. To describe\nthe low-energy physics, an effective range expansion is\nusually introduced to incorporate the energy dependence\nof coupling strength g(E), which reads\n1\ng(E)=1\ng(0)+r0E, (1)\nwhereEis the energy of two colliding particles in the\ncenter-of-mass frame, and g(0) is coupling strength at\nthreshold energy. Here r0is the effective range, which\ncrucially depends on the width of resonance. In par-\nticular, for narrow resonances, r0is typically large as\nto be comparable with the inter-particle distance. Pre-\nvious studies have revealed interesting effects of finite\nrange in cold atoms systems, including the generation\nof stronger interaction effects[ 2–6], the modification of\nFermi superfluids[ 7,8], the sub-leading high-momentum\ntail[9–14], and the stabilization of repulsive polaron[ 15]\nand p-wave system[ 16].\nIn this work, we reveal the significant effect of effective\nrangetothequantummagnetismofspin-1 /2Fermigases.\nHere we take the 1D Fermi gas in strongly repulsive\nregime, which features an impenetrable nature and thus\nsupports a hidden “lattice” structure. In this regime, the\nsystem is well described by an effective Heisenberg spin-\nchain Hamiltonian which exhibits exotic quantum mag-\nnetic properties and provides additional insights into the\nsimulation of quantum magnetism without lattice[ 17–30]\nand suchspin-chainsystemswererealizedexperimentally\nrecently[31]. Previous studies based on zero-range inter- AFMFM Mixed \nPhase \n0/10g0\nFIG. 1. (Color online). Schematic diagram of magnetic order\nofaspin-1 /2fermion systemwithafinite-range( r0/negationslash= 0) inter-\naction in a harmonic trap. As we tune the interaction (1 /g0)\nfrom the repulsive to the attractive side, the system adiaba t-\nically goes from a AFM- correlated spin chain to FM-AFM-\nFM mixed and finally to fully FM-correlated chain, which\nis in sharp contrast to the zero-range case with either AFM\nphase (1 /g0>0) or FM phase(1 /g0<0). The red and green\nbonds represent the AFM and FM couplings between spins,\nrespectively.\nactions have shown that the system can host either an\nanti-ferromagnetic (AFM) or a ferromagnetic (FM) spin\ncorrelation, depending on the sign of coupling strength\nor other small perturbations. Moreover, at the infinite\ncoupling (hard-core) limit, spin and charge are fully de-\ncoupled and the system exhibits huge spin degeneracy, at\nwhich point the FM transition is predicted[ 32]. Here, we\nwill show that the inclusion of a finite effective range can\nqualitatively change above conclusions and brings much\nricher magnetic structures to the system.\nOur results can be summarized in Fig. 1. In the pres-\nence of a finite range r0, as tuning the inverse coupling\n1/g0from the repulsive to attractive side, the system2\nadiabatically goes from a AFM-correlated spin chain to\nFM-AFM-FM mixed and finally to fully FM-correlated\nchain. The spatially modulated magnetic correlation is\ndue to the range-modified Heisenberg coupling in an ef-\nfective spin chain model, which is more sensitive to the\nlocal density as compared to that induced by bare cou-\npling (see Eq. 10). Our results demonstrate an adiabatic\nformationofFM domainsfromthe AFM state, whichhas\nnot been achieved up to date. We further propose to ver-\nify these magneticpropertiesinthe tunneling experiment\nof tilted harmonic potential.\nRange-modified effective spin chain. We begin with de-\nriving an effective spin chain model for strongly repulsive\nspin-1/2 fermions( ↑,↓) in the presence of a finite range.\nThe original Hamiltonian is H=H(0)+U, (here/planckover2pi1= 1)\nH(0)=/summationdisplay\ni/parenleftbigg\n−1\n2m∂2\n∂x2\ni+1\n2mω2\nhox2\ni/parenrightbigg\n,(2)\nU=/summationdisplay\ni,jgijδ(xi↑−xj↓). (3)\nHeregijfollows the effective range expansion: 1 /gij=\n1/g0+r0Eij, withg0the bare coupling constant and Eij\nthe relative energy of two colliding particles xi↑andxj↓.\nThepresentstudywillfocusonthenearresonanceregime\nwith large g0and small r0. Here the confinement length\nis defined as aho= (mωho)−1/2.\nTo highlight the range effect, let us first consider the\ncase ofg0=∞. In this case, without the range ( r0= 0)\nthe collision of atoms is forbidden due to hard-core in-\nteraction and the spins can distribute in an arbitrary\norder in coordinate space, giving the large spin degen-\neracy. When turn on the range ( r0/ne}ationslash= 0), however, the\natoms only experience hard-core interaction at zero rel-\native energy( Erel= 0) but not at finite Erel, and the\nfinite-Erelinteraction causes a super-exchange of spins\nat neighboring orders in the coordinate space, giving rise\nto an effective spin chain model. Following the standard\nprocedure, we obtain the effective Heisenberg spin-chain\nsolely induced by r0:\nHr\neff=r0/summationdisplay\nlJr\nl/parenleftbigg\nsl·sl+1−1\n4/parenrightbigg\n, (4)\nherelis the order index of particles in coordinate space,\nand the Heisenberg coupling is\nJr\nl=2N!\nm2/integraldisplay\ndxEij/vextendsingle/vextendsingle/vextendsingle∂D\n∂xij|xij=0/vextendsingle/vextendsingle/vextendsingle2\nθ(···< xi=xj<···),(5)\nwherexij=xi−xj;D({xi}) is the Slater determinant of\nN fermions occupying the lowest N-level of H(0), andEij\nis the relative collision energy of two particles ( xi, xj)\nin theD({xi}); in the θ-function xi(=xj) is with order\nindexl.\nTo verify the range-induced spin-chain model, we have\nexactly solved the two-body ( ↑↓) and three-body ( ↑↑↓)r0\nFIG. 2. (Color online). Energy spectrum for 1D trapped\n↑↓and↑↑↓systems as a function of effective range r0at 1D\nresonance (1 /g0= 0). The linear fit is based on the effective\nspin-chain model ( 4). Here the energy and length units are\nωhoandahorespectively.\nproblems in a trapped system with tunable range. In\nFig.2, we plot the obtained energy spectra of these sys-\ntems as a function of r0at 1/g0= 0, in comparison with\nthe linear fit from the model Hr\neff. We can see that the\neffective model can well predict the real spectra of both\nsystems for small r0(≤0.1aho). This validates the effec-\ntive spin-chain model used for larger systems.\nIn combination with the spin-chain model for small\n1/g0[18–28], we write the final effective model in the\nlimit of large g0and small r0as:\nHeff=/summationdisplay\nl/parenleftbigg1\ng0Jg\nl+r0Jr\nl/parenrightbigg/parenleftbigg\nsl·sl+1−1\n4/parenrightbigg\n,(6)\nwithJg\nl:\nJg\nl=2N!\nm2/integraldisplay\ndx/vextendsingle/vextendsingle/vextendsingle∂D\n∂xij|xij=0/vextendsingle/vextendsingle/vextendsingle2\nθ(···< xi=xj<···).(7)\nPhysically, both the bare coupling ( g0) and effective\nrange (r0) produce the same isotropic Heisenberg term\nis because both of them take effect in the spin-singlet in-\nteraction channel, thus the effective model is determined\nby the same spin-projection operator [ 27].\nHeisenberg couplings. Before studying the quantum\nmagnetism, we first examine the density dependence\nofJg\nl, Jr\nlfor a homogeneous large system, and the\ntrapped case can be deduced from the local density ap-\nproximation(LDA). Previously, Jg\nlwas shown to depend\non the cubic density ( ∼n3)[18,33], by extrapolating\nthe nearest-neighboring exchange coupling in Hubbard\nmodel to continuum[ 34,35]. Here we point out an alter-\nnative way to derive Jg\nlandJr\nlfrom Eqs.( 7,5) through3\n−3 −2 −1 0 1 2 30510 15 20 25 \n−3 −2 −1 0 1 2 30246LDA v.s Exact\n LDA v.s Exact\nglJrlJ\n2 Nl− 2 Nl−(a) (b)\nFIG. 3. (Color online). Heisenberg couplings Jg\nl(a) and\nJr\nl(b) (in the units of1\nm2a3\nhoand1\nm3a5\nhorespectively) in a\nharmonic trap. The star, circle, diamond, square, and trian -\ngular points are exact solutions of Eqs.( 8,9) for total number\nN= 2,3,4,5,6. The red lines are from analytical expressions\n(10) together with Thomas-Fermi density (see text).\nthe momentum averaging below the Fermi sea:\nJg=2n\nm2/angbracketleftBig/parenleftbiggk1−k2\n2/parenrightbigg2/angbracketrightBig\n, (8)\nJr=2n\nm3/angbracketleftBig/parenleftbiggk1−k2\n2/parenrightbigg4/angbracketrightBig\n, (9)\nhere/an}bracketle{tF(k1,k2)/an}bracketri}ht ≡/integraltext /integraltext\nF(k1,k2)dk1dk2//integraltext /integraltext\ndk1dk2, and\nthe integration is for k1,k2∈[−kF,kF], with kF\nthe Fermi momentum determined by the density n=\nk3\nF/(6π2). The essence of Eqs.( 8,9) is to reformulate the\nmany-fold integration into the combination of the local\ndensity and a pair-averaged function in terms of the rel-\native momentum of two particles within the Fermi sea.\nThe procedure leads to\nJg=2π2n3\n3m2, Jr=4π4n5\n15m3. (10)\nRemarkably, here the range-induced coupling Jrhas a\nmuchmoresensitivedependenceonthelocaldensitythan\nJg, which we will show below to significantly affect the\nquantum magnetism in the trapped system. In Fig. 3,\nwe show Eq. 10can well reproduce the exact solutions of\nJg\nl, Jr\nlfrom Eqs.( 8,9) for trapped systems up to N= 6,\nhere for the local density we have used Thomas-Fermi\napproximation nl→n(¯xl) =1\nπ/radicalBig\n2m(NωT−1\n2mω2\nT¯x2\nl),\nwith ¯xl=/an}bracketle{t(xl+xl+1)/2/an}bracketri}ht.\nTo this end we can rewrite the spin-chain Hamiltonian\n(6) asHeff=/summationtext\nlJeff\nl/parenleftbig\nsl·sl+1−1\n4/parenrightbig\n, where the effective\ncoupling depends on the coupling, range, and local den-\nsitynl:\nJeff\nl=2π2n3\nl\n3m2/parenleftbigg1\ng0+2π2\n5mn2\nlr0/parenrightbigg\n. (11)\nSpatially modulated quantum magnetism. From the ex-\npression of Jeff\nl, we can see that its sign can be effectivelytuned by local density nl, distinct from the zero-range\ncase where the magnetic property is solely determined by\nthe sign of coupling strength. This immediately leads to\ntwo effects for trapped system with inhomogeneous den-\nsity. First, the system is no longer described by a single\ncoupling strength, and there is no exact hard-core limit\nwith largespin degeneracy. Secondly, given r0>0 and in\nthe regime of 1 /g0<0, particles at different regions in-\nside the trap may experience different signs of Jeff, which\nmeans that the AFM and FM magnetic correlation can\ncoexist in the system, i.e., the quantum magnetism can\nbe locally manipulated.\n 0.25 0.2 0.15 0.1 0.05 7.85 7.9 7.95 88.05 \n1/ g0E(a)\n\u0012\n\u0013\n\u0014\n\u0015 USJQMF\u0001EFHFOFSBUF EPVCMF\u0001EFHFOFSBUF \u0001\u0001\u0001\u0001\u0001\u0001\u0001\u0001\"'. \n\u0001\u0001\u0001\u0001\u0001\u0001\u0001\u0001\u0001\u0001'. \u0012\u0013\u0014\u0015\u0016\n\u0016\nOPO\u000eEFHFOFSBUF \n@(b) \nFIG. 4. (Color online). (a) Energy spectrum of four fermions\n(N↑=N↓= 2) as a function of −1/g0(in the unit of maho) in\na trapped system. Here we take r0= 0.05aho. (b) Schematics\nof AFM (red lines) and FM (green lines) magnetic correlation\nin the chain at different coupling strengths as marked in (a).\nThe dashed lines with crossings refers to the zero effective\ncoupling Jeff\nl= 0.\nIn Fig.4(a), we show the energy spectrum of four\nfermions ( N↑=N↓= 2) by solvingthe spin-chainHamil-\ntonian (6) at a given r0= 0.05aho. Totally there are\nsix energy levels, two with total spin S= 0, three with\nS= 1 and one with S= 2, similar to the zero-range\ncase[5]. With a finite range, the six levels no longer cross\neach other at 1 /g0= 0, while the ground state transition\n(fromS= 0 toS= 2) moves to the negative coupling\nside atgc<0.\nNow we analyze the adiabatic change of the ground\nstate (with S= 0) before the transition (in the regime\n−1/g0<−1/gc). In the positive coupling side of reso-\nnance, the ground state holds the AFM correlations in\nthe spin chain given by all positive Jeff\nl(marked as 1○\nin Fig.4(a,b)). As increasing −1/g0across resonance to\nnegative coupling side, two states becomes degenerate at\npoint2○, where the Heisenberg coupling at the edges of\nthe chain touches zero( Jeff\n1=Jeff\n3= 0) and the edge par-\nticlesaredecoupledfromthechain, givingtwodegenerate\nspin states ( S= 0 and S= 1). Further increasing 1 /g0\nbeyond this point, the edge spins have FM correlation\nwith negative coupling, while the center are AFM corre-\nlated with positive coupling (see 3○). When reaching 4○,4\nthe center coupling becomes zero ( Jeff\n2= 0), and the sys-\ntem is divided into two independent magnetic domains\nwith FM correlation. At this point, three spin states are\ndegenerate, correspondingto twoFM(triplet) pairsform-\ning total spin S= 0,1,2. Beyond this point, the ground\nstate of the system changes to S= 2 FM state, and all\nsites are with FM correlations ( 5○).\n0 0.005 0.01 0.015 0.02 −0.1−0.08−0.06−0.04−0.020\nr01/g 0\nFM AFM \nMixed\nFIG. 5. (Color online). Phase diagram of magnetic or-\nder in terms of the bare coupling and effective range. Here\nN↑=N↓= 4. The units of 1 /g0andr0aremahoandaho\nrespectively.\nAbove picture can be generalized to an arbitrary num-\nber of particles. Take the spin-balanced case N↑=N↓=\nN/2 for example, as increasing −1/g0the ground state of\nthe system(with S= 0) is expected to cross a sequence\nof degeneracies with other spin states until the transition\nto FM ( S=N/2) state. The first degeneracy occurs\nat 1/g0=−2π2\n5mn2\n1r0, whenJeff\n1=Jeff\nN−1= 0 and two\nedge spins are separated from the system giving a two-\nfold degeneracy, see the upper line in Fig. 5. Increasing\n−1/g0beyond this point, the Heisenberg couplings tran-\nsit from pure AFM type to FM-AFM-FM mixed type.\nIn this mixed phase, the trap center shows the AFM\ncorrelation (with Jeff>0) because of higher density,\nwhile the trap edge shows FM correlation with Jeff>0\nbecause of lower density. In this regime, a m+ 1-fold\n(1≤m≤N/2) degeneracy occurs at 1 /g0=−2π2\n5mn2\nmr0\nwhenJeff\nm=Jeff\nN−m= 0. Continuously increasing −1/g0,\nthe regions with FM correlations becomes enlarged while\nAFM correlation becomes reduced, giving the increasing\n/an}bracketle{tS2\nLS2\nR/an}bracketri}ht, whereSL/Ris the total spin of left/right part\nof the trapped system. The ground state transition to\nFM state happens exactly at the N/2 + 1 fold degen-\neracy point(see lower phase boundary in Fig. 5), when\n1/g0=−2π2\n5mn(0)2r0withn(0) the density at trap cen-\nter. At this point/radicalbig\n/an}bracketle{tS2\nLS2\nR/an}bracketri}htreaches the maximum value\nN/4(N/4 + 1), suggesting two large and separated FMdomains formed at the left and right regions of the trap.\nTunneling experiment . Now we come to the exper-\nimental detection of the range-induced magnetic or-\nders, using the tunneling techniques as established in\nexperiments[ 31]. By varying the magnetic field gradi-\nent and tilting the potential barrier, one can control the\nnumber of atoms tunneling out of the trap and probe the\nspin structure. Here we propose the measure the possi-\nbility of having all spin- ↓atoms tunneling from the right\nside of the trap, which is given by the weight of the full\nspin separated configuration ( | ↑↑...↓↓.../an}bracketri}ht) in the wave\nfunction, as denoted by P↓in following discussions.\n 0.1 0.05 0 −0.05 −0.100.050.10.150.20.250.30.33\n \n4 5 6 7 8 9 10 00.05 0.1 0.15 0.2 0.25 0.3 0.33 \n P\nP\n-1/g 0 NMixed(r0=0.02)\nFM(r0=0)\nAFM(r0=0)(a) (b) Mixed \nFM \nAFM \nFIG. 6. (Color online). (a) Probabilities of all spin- ↓atoms\ntunneling from the right side of the titled trap, denoted by\nP↓. Red solid line is for system with range r0= 0.02aho, by\nadiabatically following the ground state in g0>0 side to g0<\n0 side. The cross point marks the location of FM transition,\nwhenP↓shows a maximum. Blue dashed and green dotted\nlines are for zero-range systems adiabatically following t he\nAFM ground state( S= 0) and the FM state( S=N/2). Here\nN↑=N↓= 2, the unit of 1 /g0ismaho. (b) The maximum\nofP↓as a function of particle number Nfor spin-balanced\nfermions. Red, blue and green lines are the same as in (a).\nIn Fig.6(a), we show P↓as a function of −1/g0for\nN↑=N↓= 2 system, by adiabatically following certain\nstatefromthe g0>0side. Without range,weseethat P↓\nis very small ( ∼4%) if following the AFM ground state,\nandcanbeaslargeas16 .7%iffollowingFMstate, consis-\ntent with the experiment [ 31]. Turning on the range and\nfollowing the ground state in g0>0 side,P↓is no longer\na constant but varies sensitively with −1/g0, suggesting\nthesignificantchangeofmagneticstructures/correlations\nin the trap. In particular, we see that P↓reaches a max-\nimum near the N/2+1-fold degeneracy point, where the\neffective coupling at the trap center touches zero and the\nsystem is composed of two FM domains (each with spin\nS=N/4). The maximum value can be then deduced by\nexpanding the S= 0 state by two spins with S=N/4:\n|S= 0,Sz= 0/an}bracketri}ht=N\n4/summationdisplay\nm=−N\n4C(N)\nm|S1=N\n4,m/an}bracketri}htN\n2|S2=N\n4,−m/an}bracketri}htN\n2,\n(12)5\nthenP↓is exactly given by the Clebsch-Gordor coeffi-\ncients as:\nP↓=/vextendsingle/vextendsingle/vextendsingleC(N)\nN\n4/vextendsingle/vextendsingle/vextendsingle2\n=1\nN/2+1. (13)\nIn Fig.6(b), we have verified this analytic result by nu-\nmerically calculations from the spin-chain model. Re-\nmarkably, Eq. 13produces a much larger P↓in compar-\nison to the FM state, where P↓=N\n2!N\n2!\nN!≃√\nπN\n2\n2Nis\nexponentially small. This reflect the distinct magnetic\nstructure of a pure FM state and a composition of two\nFM domains. This can serve as experimental evidence\nto identify the spatially-modulated quantum magnetism\ndue to the finite range effect.\nFinal remark. Our results reveal the significant effect\nof a finite effective range in the strong coupling regime\nof 1D trapped spin-1 /2 fermions. The sensitive density-\ndependence of Heisenberg coupling induced by the finite\nrange suggests a convenient route towards the local ma-\nnipulation of quantum magnetism. In particular, by en-\ngineering the density distribution of cold atomic gases\nthrough the laser potentials, one may get access to an ar-\nbitrary configuration of local Heisenberg coupling in the\ncoordinate space, and thus an arbitrary type of magnetic\norder may be achieved. This concept can be generalized\nto higher spins and other composition ofatomic mixtures\nin the 1D strong coupling regime.\nAcknowledgment. The work is supported by the Na-\ntional Key Research and Development Programof China\n(2018YFA0307600, 2016YFA0300603), and the National\nNatural Science Foundation of China (No.11622436,\nNo.11425419, No.11421092, No.11534014).\n∗xlcui@iphy.ac.cn\n†schen@iphy.ac.cn\n[1] C. Chin, R. Grimm, P. Julienne, and E. Teisinga, Rev.\nMod. Phys. 82, 1225 (2010).\n[2] T.-L. Ho, X. Cui, W. Li, Phys. Rev. Lett. 108, 250401\n(2012).\n[3] E. L. Hazlett et al, Phys. Rev. Lett. 108, 045304 (2012).\n[4] R. Qi and H. Zhai, Phys. Rev. A 85, 041603(R) (2012).\n[5] X. Cui, Phy. Rev. A 86, 012705 (2012).\n[6] R. Qi and X. Guan, Europhysics Letters 101, 40002\n(2013).[7] V. Gurarie and L. Radzihovsky, Annals of Physics 322,\n2 (2007).\n[8] Y. Nishida, Phys. Rev. Lett. 109, 240401 (2012)\n[9] S. B. Emmons, D. Kang and L. Platter, Phys. Rev. A 94,\n043615 (2016).\n[10] Z. Yu, J. H. Thywissen, S. Zhang, Phys. Rev. Lett. 115,\n135304 (2015); see also Erratum: Phys. Rev. Lett. 117,\n019901 (2016).\n[11] C. Luciuk, S. Trotzky, S. Smale, Z. Yu, S. Zhang, and J.\nH. Thywissen, Nat. Phys. 12, 599 (2016).\n[12] M.-Y. He, S.-L. Zhang, H. M. Chan, Q. Zhou, Phys. Rev.\nLett.116, 045301 (2016).\n[13] X. Cui and H. Dong, Phys. Rev. A 94, 063650 (2016)\n[14] X. Yin, X.-W. Guan, Y. Zhang, H. Su, and S. Zhang,\nPhys. Rev. A 98, 023605 (2018).\n[15] C. Kohstall et al, Nature 485, 615 (2012).\n[16] L. Pan, S. Chen, X. Cui, Phys. Rev. A 98, 011603 (R)\n(2018).\n[17] A. G. Volosniev, D. V. Fedorov, A. S. Jensen, M. Va-\nliente, and N. T. Zinner, NatureCommunications 5, 5300\n(2014).\n[18] F. Deuretzbacher, D. Becker, J. Bjerlin, S. M. Reimann,\nand L. Santos, Phys. Rev. A 90, 013611 (2014).\n[19] J. Levinsen, P. Massignan, G. M. Bruun, M. M. Parish,\nScience Advances 1, e1500197 (2015)\n[20] A. G. Volosniev, D. Petrosyan, M. Valiente, D. V. Fe-\ndorov, A. S. Jensen, and N. T. Zinner, Phys. Rev. A 91,\n023620 (2015).\n[21] L. Yang, L. Guan, and H. Pu, Phys. Rev. A 91, 043634\n(2015).\n[22] P. Massignan, J. Levinsen, and M. M. Parish, Phys. Rev.\nLett.115, 247202 (2015).\n[23] H. Hu, L. Guan, and S. Chen, New J. Phys. 18, 025009\n(2016).\n[24] H. Hu, L. Pan, and S. Chen, Phys. Rev. A 93, 033636\n(2016).\n[25] L. Yang and H. Pu, Phys. Rev. A 94, 033614 (2016).\n[26] F. DeuretzbacherandL.Santos, Phys.Rev.A 96, 013629\n(2016).\n[27] L. Yang and X. Cui, Phys. Rev. A 93, 013617 (2016).\n[28] L. Yang, X.-W. Guan, X. Cui, Phys. Rev. A 93, 051605\n(R) (2016).\n[29] F. Deuretzbacher, D. Becker, J. Bjerlin, S. M. Reimann,\nand L. Santos, Phys. Rev. A 95, 043630 (2017).\n[30] L. Pan, Y. Liu, H. Hu, Y. Zhang, and S. Chen, Phys.\nRev. B96, 075149 (2017).\n[31] S. Murmann, F. Deuretzbacher, G. Z¨ urn, J. Bjerlin, S.\nM. Reimann, L. Santos, T. Lompe, S. Jochim, Phys. Rev.\nLett. 115, 215301 (2015).\n[32] X. Cui and T.-L. Ho, Phys. Rev. A 89, 023611 (2014).\n[33] O. V. Marchukov, E. H. Eriksen, J. M. Midtgaard, A. A.\nS. Kalaee, D. V. Fedorov, A. S. Jensen, and N. T. Zinner,\nEur. Phys. J. D 70, 32 (2016).\n[34] M. Ogata and H. Shiba, Phys. Rev. B 41, 2326 (1990).\n[35] K. A. Matveev, Phys. Rev. B 70, 245319 (2004)." }, { "title": "0802.3646v1.Tunnel_barrier_enhanced_voltage_signals_generated_by_magnetization_precession_of_a_single_ferromagnetic_layer.pdf", "content": " 1 Tunnel barrier enhanced voltage signals generate d by magnetization precession of a single \nferromagnetic layer \n \nT. Moriyama,1 R. Cao,1 X. Fan,1 G. Xuan,2 B. K. Nikoli ć,1 Y . Tserkovnyak,3 J. Kolodzey,2 and \nJohn Q. Xiao1 \n1Department of Physics and Astronomy, Univer sity of Delaware, Newark, DE 19716, USA \n2Department of Electrical and Computer Engineeri ng, University of Delaware, Newark, DE 19716, USA \n3Department of Physics and Astronomy, University of California, Los Angeles, CA 90095, USA \n \n \n \nAbstract: \nWe report the electrical detection of magnetization dynamics in an Al/AlO x/Ni80Fe20/Cu tunnel \njunction, where a Ni 80Fe20 ferromagnetic layer is brought into precession under the \nferromagnetic resonance (FMR) conditions. The dc voltage generated across the junction by the \nprecessing ferromagnet is enhanced about an order of magnitude compared to the voltage signal \nobserved when the contacts in this type of mu ltilayered structure are ohmic. We discuss the \nrelation of this phenomenon to magnetic spin pumping and speculate on other possible \nunderlying mechanisms responsible for the enhanced electrical signal. \n \nPACS numbers: 76.50.+g, 72.25.Mk, 72.25.Hg \n 2 In recent years, basic and applied research in metal-based spintronics has shifted \nincreasingly from the static to the dynamic magnetic properties in hybrid nanostructures \ncomposed of ferromagnetic and normal metal layers [1-5]. A variety of experimentally observed \nphenomena involving nonlocal magnetization dynamics in magnetic multilayers are due to two \ncomplementary effects: (i) the transfer of spin angular momentum accompanying charge \ncurrents driven by the applied bias voltage between ferromagnetic layers results in torques that \n(for sufficiently high current densities) ge nerate spontaneous magnetization precession and \nswitching [1]; and (ii) the precessi ng magnetization of a ferromagnet ( FM) pumps spins into \nadjacent normal metal layers ( NM) with no applied bias [2, 5, 6]. In particular, the spin pumping \neffect [5] is a promising candidate for realizing a spin battery device [7] as a source of elusive \npure spin currents (not accompanied by any net charge transport) emitted at the FM/NM \ninterface, where steady magnetization precession of the FM layer is sustained by the absorption \nof external rf radiation under the FMR conditions. Another promising application of \nmagnetization dynamics is microwave-assist ed reduction of the switching field of FM, which \ncould play an important role in advanced recording media [4]. \nThus far, however, the spin pumping effect has been demonstrated mostly indirectly as \nan additional contribution to the FMR linewidth in FM/NM multilayers (where the NM is Pt, Pd, \nCu, etc.) that can be described as the inte rface-induced enhancement of the Gilbert dumping \nconstant [8-11]. The vigorous pursuit of direct el ectrical detection of spin pumping has led to \ntheoretical proposals [6] to use a single precessing FM as both the source and detector of \npumped spin accumulation in NM layers. This adroit scheme has b een realized in a very recent \nexperiment [2] measuring the diff erence in voltages of the orde r of several hu ndred nanovolts 3 between two FM/NM interfaces of a NM1/FM/NM2 trilayer. \nIn this Letter, we report measuremen ts of the dc voltage across Al/AlO x/Ni80Fe20/Cu \ntunnel junctions with precessing magnetization of Ni 80Fe20. The surprisingly large observed \nvoltage is about μV , which seems to qualitatively agree with the spin pumping theory [5, 6] but \nrequires an unreasonably large spin-mixing conductance of the FM/tunnel-barrier contact. We \nconclude that a new nonequilibrium phenome non, which dynamically couples the spin and \ncharge degrees of freedom, exists in tunneling stru ctures. The rest of this Letter presents details \nof our experiment, and we conclude by speculati ng on possible theoretical scenarios responsible \nfor these surprising experimental results. \nThe experimental setup is illustrated in Fi g. 1. A tunnel junction was fabricated on the \nsignal conductor of a coplanar waveguide (C PW) transmission line. The tunnel junction \nstructure of Cu (100nm)/Al (10nm)/AlO x (2.3nm)/Ni 80Fe20 (20nm)/Cu (70nm)/Au (25nm) \nwas fabricated on a Si substrate with a 1 μm thick thermal oxide layer, by using magnetron \nsputtering deposition and conventional microfab rication processing. The bottom-most 100 nm \nCu layer was patterned into the CPW designed to have 50 Ω characteristic impedance in the \nabsence of the tunnel structure. The aluminum oxide tunnel barrier was formed by plasma \noxidation. The size of the tunnel junction pillar was 50×50 μm2, and the dc junction resistance \nwas measured to be 67 k Ω. A microwave signal from a vect or network analyzer (Agilent \n8753B) was introduced through the CPW and generated a microwave magnetic field rfHthat \nwas linearly polarized in the plane of the tunne l junction. The external dc magnetic field (-120 \nOe to 120 Oe) was swept along the axis of the CPW (the y-direction), so that the magnetization \nchanged its direction within the x-y plane. The precessing spin mainly rotated around the y-axis. 4 Two electrical probe tips were used to measure the dc voltage across the junction. The \nmicrowave input signal was varied from 0.7 GHz to 4 GHz, with power up to 18 dBm \namplitude-modulated with a 400 Hz sinusoidal signal to allow for lock-in detection. It should be \nmentioned that there was always a few tens of microvolt background voltage at the detector due \nto microwave noise. We found th at the background voltage was dire ctly proportional to the input \nmicrowave power. We can thus maintain the sa me power applied to the device at different \nfrequencies by slightly tuning the nominal input power (±1 dB m) to maintain the same \nbackground voltage within 20% error. This ma kes it possible to compare the data between \ndifferent frequencies without concerning the frequency dependence of the CPW impedance \nwhich changes the power input to the device slightly. \nFigure 2 shows the dc voltage as a functi on of the external magnetic field in the \nAl/AlO x/Ni80Fe20/Cu tunnel junction. At each microwav e frequency, the voltage peaks of \nmagnitude VΔappear symmetrically at positive and negative fields. The peak field as a \nfunction of the microwave frequency shown in Fi g. 3 agrees well with the values we obtained \nfrom the flip-chip CPW FMR measurements. Th e Kittel formula [12] fits the data with \nreasonable parameters, 4 9 kGsMπ= , 19 OekH= , and gyromagnetic ratio \n-1 -10.0176 s Oeγ= , confirming that the dc voltage peak appears at the uniform FMR mode of \nthe Ni 80Fe20 layer. The peak magnitude reaches about 1 μV at 2GHz which is much larger than \nthe maximum value of about 250 nV at 14.5 GHz reported in Re f. [2] for a Pt/Ni 80Fe20/Al \nstructure. Figure 4 shows microwav e power and frequency dependence of VΔ, which increases \nwith increasing microwave power [Fig. 4 (a)]. We also plot VΔ as a function of precession \ncone angle in Fig. 4(b). Th e precession cone angle of Ni 80Fe20 was determined by the change in 5 the tunnel resistance at the FMR field in IrMn/Fe 70Co30/AlO x/Ni80Fe20 magnetic tunnel junctions \n[13] with 20mV bias voltage so that the dc voltage effect of the microvolt order we are \ndiscussing here can be neglected. A clear dip in an tiparallel states and a pe ak in parallel states \nare observed corresponding to FMR fields, and the precession angle θ can be determined from \n()1c o s RR θ Δ∝ − . At 10 dBm power input, the precessi on cone angle was around 17° [13]. \nAs the applied frequency increased, VΔ increases almost linearly, as shown in Fig. 4(c). \nBefore attempting to interpret our results, we have to examine carefully the \nrectification effects which could induce similar dc voltage response. Possible rectification effect \nmay arise from both the time-dependent anis otropy magnetoresistance (AMR) effect and the \nanomalous Hall effect (AHE) discussed in Refs.[ 14]. The current due to these two effects is \ngiven by \n()2RMρσρΔ′=− ⋅ + ×jj MM jM, (1) \nwhere j is the current, σ is the conductivity, ρ is the resistivity, ρΔ is the \nmagnetoresistive anisotropy, and R is the anomalous Hall constant. The magnetization \nprecessing around the y-axis is described by the vector \n()s i ns i n ,c o s ,c o ss i nMt M M tωθθω θ =M . The microwave-induced current across the \njunction along the microwave electric fi eld direction (z-axis) is given by \n()() αω+ = t jcos,0,0j , where αis the phase difference with respect to the phase of spin \nprecession . The z -component of Eq. (1) is of in terest to our experiments, \n() θωαωρρ2 2sin cos cos ' t t j jz + Δ−= . The time average of zj′ is zero, allowing us to \nconclude that there is no dc component in the z -direction in our sample. The result holds even if 6 the precession axis fluctuates in the x-y plane. Thus, in our sample c onfiguration we expect no \ndc voltage generated due to the rectification effect. We purposely broke the tunnel barrier to \ninvestigate Al/Ni 80Fe20/Cu contact and also made Cu/AlO x/Al/Ni 80Fe20/Cu junctions. In both \ncases, no VΔ was observed within our measurement se nsitivity of about 100 nV . This implies \nthat the large VΔ in Al/AlO x/Ni80Fe20/Cu was indeed developed due to the AlO x/Ni80Fe20 \ninterface. \nLet us now try to interpret our results w ithin the framework of the standard spin \npumping theory [5-7]. At the FMR, a steady precession of the magneti zation of the FM layer \npumps a spin current into the adjacent NM according to [5, 7] \nRe Im4pump\nsddggdt dt π↑↓ ↑↓⎛⎞=× +⎜⎟⎝⎠mmIm=, (2) \nwhere m is the unit vector along th e instantaneous direction of the precessing magnetization \nand Reg↑↓ (Img↑↓) is the real (imaginary) part of the dimensionless interfacial spin-mixing \nconductance (in units of 2eh ) which describes spin tr ansport perpendicular to mat the \nFM/NM interface [5, 15, 16]. For transparent intermetallic FM/NM contacts Img↑↓ is \ntypically neglected because of being much smaller than Reg↑↓[5-7], while for low transparent \ncontacts we find Re 0.5gg↑↓= and Im 0.5gg↑↓\u0011 (using the simple Stoner model for \nFM and random binary alloy w ith a gap in the energy spectru m for the tunnel barrier as NM \nlayer; g is the total charge conductanc e of the junction). The possi bility for non-negligible \nImg↑↓ for tunneling interfaces is also highlighted by recent first principles calculations [17]. \nThe injected spin current builds up a spin accumulation sμ in the NM layer (close to the \nFM/NM interface) when the spin -flip relaxation rate in NM is smaller than the spin injection rate. 7 This, in turn, drives a backward flowing spin current back\nsIinto the precessing FM [5]. The \nbackward flowing spin current parallel to the magnetization can be absorbed by the FM, in the \npresence of spin-flip processes . Due to spin-dependent bulk and interface conductances, this \nabsorbed spin current is converted into charge accumulation at the FM/NM interface [18]. The \nmaximum value of the voltage drop dcVacross the FM/NM interface, at fixed frequency ω \nand cone angle θof magnetization precession (assuming the frequency is much greater than the \ncharacteristic spin-flip rate in the normal me tal and the ferromagnet is thicker than its bulk \nspin-diffusion length), is obtained for NM layer thickness much sma ller than its spin-diffusion \nlength dN << λdsN as [6]: \n()2\ndc 22 2 2sin cos\n2 () s i n ( 1 ) c o sF\nFN FpgVe gp gg p g g gω\nωω ω ω ω ωθθ ω\nηθ θ↑↓=\n−+ + + −=, (3) \nwhere ()tanhN\nNN N s dgg dω η λ↑↓= with the thickness dN and the spin diffusionN\nsdλ of the \nNM layer, Fg and Ng parameterize the transport properties of the bulk FM and NM, gω↑↓ \nis the real part of the eff ective spin-mixing conductance, ()( ) p gg ggω ωω ωω↑↓↑ ↓=− + is the \ninterfacial spin-polarization, gggωωω↑↓=+ is the sum of spin-up and spin-down effective \nconductances of the FM/NM interface. The “ef fective” interfacial transport quantities are in \ngeneral frequency-dependent since they have to be evaluated for the interface resistance in series \nwith a NM resistor of length NM LDω ω= over which the oscillati ng transverse components \nof spin accumulation in NM (with diffusion constantNMD) are averaged to zero, although this is \nnot important in practice for high-impeda nce tunnel barriers. The voltage drops 1,2\ndcVemerging \nat each of the two FM/NM contacts will differ from each other when conductances gω↑↓ and/or \nspin-flip diffusion lengthsN\nsλon two sides of the multilayer ar e substantially different, as 8 observed by measuring 12\ndc dc VV VΔ= − on lateral Pt/Ni 80Fe20/Al device in a recent experiment \n[2]. \nTo compare the spin pumping theory with our results, we assume that spin-mixing \nconductance is governed by the AlO x/Ni80Fe20 interface, but there is no spin flipping inside the \nbarrier and spin accumulation is induced in the Al layer. Since the interface conductance gω \nfor the low transparency tunnel barri er is much smaller than smallFg, the first term in the \ndenominator of Eq. (3) reduces to ()2sinFNgηθ+ , so that Eq. (3) can be parameterized \nwithggωω↑↓,Nη,pω, and θ. Using 0.3 pω= for the AlO x/Ni80Fe20 interface, Fig. 4(b) shows \nthe best fitting (solid line) of our results by Eq. (3), where we extract 3.4 ggωω↑↓≈ and \n0Nη≈ from the fit. We found that Nη has to be set to zero in order to fit our data, which \nrequires that Ng is roughly comparable or smaller than gω↑↓, while we expect the former to be \nseveral orders of magnitude larger than the latter, using the measured tunneling conductance. \nThis is the first discrepancy between our results and an attempt to explai n them using standard \ninterfacial spin pumping theory originally devel oped [5, 7] and experimentally confirmed [2] for \nintermetallic FM/NM contacts. On the other hand, linear fitting of VΔvs. frequency in Fig. 4(c) \nyields the slope of 0.85 μV/GHz, and 6.1 ggωω↑↓≈ at precession cone angle of 17° by Eq. (3). \nThese values are larger than the typical value 1 ggωω↑↓≈ for transparent intermetallic contacts, \nwhich is highly unexpected when compared to standard estimates [16] of ggωω↑↓ for trivial \n(non-magnetic) tunnel contacts. \nV oltage generation based on the spin pumping mechanism [6] is based on spin injection \ninto the normal metal, across the FM/NM interface, with its subsequent diffusion, relaxation, and \nbackflow into the ferromagnet, which is ultimately responsible for the build-up of the voltage 9 drop across the contact. A tunnel barrier exponen tially impedes electron flows (and thus spin \ncurrents) across the FM/NM contact, and one, therefore, would not expect a significant voltage \ngeneration by the spin-pumping mechanism. This is the reason why we were not able to reach a \nquantitative agreement with the theory. The tunne l barrier essentially cuts off the normal metal \nfrom the FM, while a voltage probe may now be thought of as a nonintrusive probe of dynamic \nprocesses within the FM. If the magnetization dynamics can generate nonequilibrium spin \naccumulation inside the ferromagnet, in analogy with the pumped spin generation in the normal \nmetal (presumably requiring spin-orbit or other spin-flip processes in the FM), the voltage \nmeasured by the FM may in fact be probing this spin accumulation rather than a nonlocal spin \npumping process. Exploring this possibility re quires further theoretical analysis and other \nnonlocal probes of the magnetization dynamics. Fi nally, we note that our theoretical discussion \ncompletely disregarded many-body effects due to electron-electron inte ractions, which may \nmodify substantially the predictions of the standa rd spin pumping theory, especially if we drive \nthe magnetization dynamics beyond the linearized regime. \nIn conclusion, we observed a large dc voltage , of the order of microvolts, across the \nAl/AlO x/Ni80Fe20/Cu tunnel junctions, due to a dynamic spin and charge coupling driven by the \nprecessing magnetization of a single Ni 80Fe20 ferromagnetic layer at ferromagnetic resonance. \nBy short circuiting the tunnel ba rrier, we demonstrated that the observed dc voltage mainly \narises from the Al/AlO x/Ni80Fe20 contact. The phenomenon appears qualitatively similar to the \npredictions of the spin pumping formalism, but a quantitative analysis shows a number of \ndiscrepancies with the standard theory. This suggests a new nonequilibrium mechanism for the \nspin and charge coupling, which is responsible fo r the voltage generation much larger than that 10 observed very recently for intermetallic interfaces [2]. We speculate on the role of intrinsic \ndynamic processes in the ferromagnet and the effect s of the electron-electr on interactions, as \npossible culprits for our observations, but a more t horough theoretical analysis is desirable in the \nfuture. \nWe thank M. D. Stiles and S.-T. Chui fo r illuminating discussions. This work was \nsupported by NSF DMR Grant No. 0405136, and DOE DE-FG02-07ER46374. 11 \nReferences: \n[1] S. I. Kiselev, et al., Nature 425, 380 (2003). \n[2] M. V . Costache, et al., Phys. Rev. Lett. 97, 216603 (2006). \n[3] J. Grollier, et al., Phys. Rev. B 73, 060409 (2006). \n[4] T. Moriyama, et al., Appl. Phys. Lett. 90, 152503 (2007). \n[5] Y . Tserkovnyak, et al., Rev. Modern Phys. 77, 1375 (2005). \n[6] X. H. Wang, et al., Phys. Rev. Lett. 97, 216602 (2006). \n[7] A. Brataas, et al., Phys. Rev. B 66, 060404 (2002). \n[8] S. Mizukami, et al., J. Magn. Magn. Mater. 226, 1640 (2001). \n[9] B. Heinrich, et al., Phys. Rev. Lett. 90, 187601 (2003). \n[10] B. Heinrich, et al., J. Supercond. Novel Magn. 20, 83 (2007). \n[11] T. Gerrits, et al., J. Appl. Phys. 99, 023901 (2006). \n[12] C. Kittel, Introduction to solid state physics (Wiley, Hoboken, NJ, 2005), 8th ed., Chap. 13. \n[13] T. Moriyama, et al., unpublished. \n[14] W. G. Egan, et al., J. Appl. Phys. 34, 1477 (1963). \n[15] K. Xia, et al., Phys. Rev. B 65, 220401 (2002). \n[16] A. Brataas, et al., Phys. Rep. 427, 157 (2006). \n[17] I. Turek, et al., J. Phys.: Condens. Matter 19, 365203 (2007). \n[18] S. Takahashi, et al., Phys. Rev. B 67, 052409 (2003). \n \n 12 Figure captions: \nFIG. 1. (Color online) Schema tic diagram of the sample stru cture (a) and the measurement \ngeometry (b). The arrow in the Ni 80Fe20 layer indicates the magnetiza tion direction. An external \ndc field exH is applied in the y-direc tion and the rf magnetic field rfH is applied along the \nx-direction. A coplanar microwave probe f eeds microwave signals through the coplanar \nwaveguide. DC voltage across the junction (TJ) is measured between top of the junction and \nsignal line of the CPW. \n \nFIG. 2. (Color online) The dc voltage VΔ generated across the Al/AlO x/Ni80Fe20/Cu tunnel \njunction as a function of the ex ternally applied static magnetic field. The frequency of the \napplied rf field ranges from 1.8 to 2.8 GH z. The background voltage is subtracted for \ncomparison purpose. \n \nFIG. 3. (Color online) The frequency dependence of static magnetic field at which the dc voltage \npeak (circles) appears in Fig. 2. The crosses label the frequency depe ndence of the resonance \nfield obtained from FMR measurem ents on flip-chip structures in the CPW line. The curve is a \nfit to the Kittel formula [12]. \n \nFIG. 4. The amplitude of the dc voltage VΔ measured across Al/AlO x/Ni80Fe20/Cu device as \nfunction of: (a) microwave power; (b) precession cone angle; and (c) microwave frequency at 10 \ndBm. Solid lines in (b) and (c) are the f it to Eq. (3) as described in the text. \n 13 \n \n \nFig.1 T. Moriyama et al. 14 \n \nFig. 2 T. Moriyama et al. 15 \n \nFig.3 T. Moriyama et al. 16 \n \nFig. 4 T. Moriyama et al. \n " }, { "title": "0810.4286v3.Magnetic_moment_manipulation_by_a_Josephson_current.pdf", "content": "arXiv:0810.4286v3 [cond-mat.supr-con] 8 Jan 2009Magnetic moment manipulation by a Josephson current\nF. Konschelle∗and A. B uzdin†\nCondensed Matter Theory Group, CPMOH, Universit´ e de Borde aux and CNRS. F-33405 Talence, France\n(Dated: November 2, 2018, Document published in Phys. Rev. L ett.102, 017001 (2009).)\nWe consider a Josephson junction where the weak-link is form ed by a non-centrosymmetric fer-\nromagnet. In such a junction, the superconducting current a cts as a direct driving force on the\nmagnetic moment. We show that the a.c. Josephson effect gener ates a magnetic precession provid-\ning then a feedback to the current. Magnetic dynamics result in several anomalies of current-phase\nrelations (second harmonic, dissipative current) which ar e strongly enhanced near the ferromagnetic\nresonance frequency.\nMany interesting phenomena have been observed re-\ncently in the field ofspintronics: the spin-dependent elec-\ntric current and inversely the current-dependent magne-\ntization orientation (see for example [1, 2]). Moreover,\nit is well known that spin-orbit interaction may be of\nprimary importance for spintronic, namely for systems\nusing a two-dimensional electron gas [3]. In the super-\nconductor/ferromagnet/superconductor(S/F/S)Joseph-\nson junctions, the spin-orbit interaction in a ferromagnet\nwithout inversion symmetry provides a mechanism for a\ndirect (linear) coupling between the magnetic moment\nand the superconducting current [4]. Similar anomalous\nproperties have been predicted for Josephson junctions\nwith spin-polarized quantum point contact in a two di-\nmensional electron gas [5]. S/F/S junctions are known\nto reveal a transition to π-phase, where the supercon-\nducting phase difference ϕin the ground state is equal\ntoπ[6]. However, the current-phase relation (CPR) in\nsuchaπ-junction hasausualsinusoidalform I=Icsinϕ,\nwhere the critical current Icdepends in a damped oscilla-\ntorymanneron themodulus ofthe ferromagnetexchange\nfield. In a non-centrosymmetric ferromagnetic junction,\ncalled hereafter ϕ0-junction, the time reversal symme-\ntry is broken and the CPR becomes I=Icsin(ϕ−ϕ0),\nwhere the phase shift ϕ0is proportional to the magnetic\nmoment perpendicular to the gradient of the asymmetric\nspin-orbit potential [4]. Therefore, manipulation of the\ninternal magnetic moment can be achieved via the super-\nconducting phase difference ( i.e.by Josephson current).\nIn the present work we study theoretically the spin dy-\nnamics associated with such ϕ0-junctions. Though there\nis a lot of experimental progress in studying the static\nproperties of S/F/S junctions, little is known about the\nspin-dynamics in S/F systems. Note here the pioneering\nwork [7] where a sharpening of the ferromagnetic reso-\nnancewasobservedbelowthe superconductingtransition\nin Nb/Ni 80Fe20system. Theoretically, the singlespin dy-\nnamicsinterplaywithaJosephsoneffecthasbeenstudied\nin [8, 9, 10, 11]. More recently, the dynamically induced\ntriplet proximity effect in S/F/S junctions was studied in\n[12, 13], while the junctions with composite regions (in-\ncluding several F regions with different magnetization)\nwere discussed in [14, 15]. Here we consider a simple\nS/F/Sϕ0-junction in a low frequency regime ℏωJ≪Tc(ωJ= 2eV/ℏbeing the Josephson angular frequency\n[16]), which allows us to use the quasi-static approach\nto treat the superconducting subsystem in contrast with\nthe case analyzed in [12, 13]. We demonstrate that a\nd.c. superconducting current may produce a strong ori-\nentation effect on the F layer magnetic moment. More\ninterestingly, the a.c. Josephson effect, i.e.applying a\nd.c. voltage Vto theϕ0-junction, would produce cur-\nrent oscillations and consequently magnetic precession.\nThis precession may be monitored by the appearance of\nhigher harmonics in CPR as well as a d.c. component of\nthe current. In particular regimes, a total reversal of the\nmagnetization could be observed. In the case of strong\ncoupling between magnetic and superconducting subsys-\ntems, complicated non-linear dynamic regimes emerge.\nTo demonstrate the unusual properties of the ϕ0-\njunction, we consider the case of an easy-axis magnetic\nanisotropy of the F material (see Fig.1). Both the easy\naxis and gradient of the asymmetric spin-orbit potential\nnare along the z-axis. Note that suitable candidates for\nthe F interlayer may be MnSi or FeGe. In these systems,\nthe lack of inversion center comes from the crystalline\nstructure, but the origin of broken-inversion symmetry\nmay also be extrinsic, like in a situation near the surface\nofathin Ffilm. Inthefollowing, wecompletelydisregard\nthe magnetic induction. Indeed the magnetic induction\nin the (xy) plane is negligible for the thin F layer consid-\nered in this paper, whereas the demagnetization factor\ncancels the internal induction along the z-axis (N= 1).\nThe coupling between F and S subsystems due to the or-\nbital effect has been studied in [17] and it appears to be\nvery weak, and quadratic over magnetic moment Mfor\nthe case when the flux of Mthrough the F layer is small\nin comparison with flux quantum Φ 0=h/2e.\nThe superconducting part of the energy of a ϕ0-\njunction is\nEs(ϕ,ϕ0) =EJ[1−cos(ϕ−ϕ0)], (1)\nwhereEJ= Φ0Ic/2πis the Josephson energy, Icis the\ncritical current and ϕ0is proportional to the Mycom-\nponent of the magnetic moment (see Fig.1). Therefore,\nwhen the magnetic moment is oriented along the z-axis,\nwe have the usual Josephson junction with ϕ0= 0. As-\nsuming the ballistic limit we may estimate the charac-2/BY\n/CB /CB\n/DE /BN /D2/DD/DC\n/C5/DD\n/B4 /D8 /B5\n/C5/DE/BDFIG. 1: Geometry of the considered ϕ0-junction. The ferro-\nmagnetic easy-axis is directed along the z-axis, which is also\nthe direction nof the gradient of thespin-orbit potential. The\nmagnetization component Myis coupled with Josephson cur-\nrent through the phase shift term ϕ0∝n.(M∧∇Ψ), where\nΨ is the superconducting order parameter ( ∇Ψ is along the\nx-axis in the system considered here).\nteristic Josephson energy as [6] Φ 0Ic/S∼Tck2\nFsinℓ/ℓ\nwithℓ= 4hL/ℏvF, whereS,Landhare the section, the\nlength and the exchange field of the F layer, respectively.\nThe phase shift is\nϕ0=ℓvso\nvFMy\nM0(2)\nwhere the parameter vso/vFcharacterizes the relative\nstrength of the spin-orbit interaction [4]. Further on we\nassume that vso/vF∼0.1. If the temperature is well be-\nlowthe Curietemperature, M0=/bardblM/bardblcanbe considered\nasaconstantequaltothesaturationmagnetizationofthe\nF layer. The magnetic energy contribution is reduced to\nthe anisotropy energy\nEM=−KV\n2/parenleftbiggMz\nM0/parenrightbigg2\n, (3)\nwhereKis an anisotropy constant and Vis the volume\nof the F layer.\nNaturally, we may expect that the most interesting\nsituation corresponds to the case when the magnetic\nanisotropy energy does not exceed too much the Joseph-\nson energy. From the measurements [18] on permal-\nloy with very weak anisotropy, we may estimate K∼\n4.10−5K.˚A−3. On the other hand, typical value of L\nin S/F/S junction is L∼10nm and sin ℓ/ℓ∼1. Then,\nthe ratio of the Josephson over magnetic energy would\nbeEJ/EM∼100 forTc∼10K. Naturally, in the more\nrealistic case of stronger anisotropy this ratio would be\nsmaller but it is plausible to expect a great variety of\nregimes from EJ/EM≪1 toEJ/EM≫1.\nLet us now consider the case when a constant current\nI < Icis applied to the ϕ0-junction. The total energy is\n(see,e.g.[16]):\nEtot=−Φ0\n2πϕI+Es(ϕ,ϕ0)+EM(ϕ0),(4)and both the superconducting phase shift difference ϕ\nand the rotation of the magnetic moment My=M0sinθ\n(whereθisthe anglebetweenthe z-axisandthe direction\nofM) are determined from the energy minimum condi-\ntions∂ϕEtot=∂ϕ0Etot= 0. It results in\nsinθ=I\nIcΓ with Γ =EJ\nKVℓvso\nvF, (5)\nwhich signifies that a superconducting current provokes\nthe rotation of the magnetic moment Myin the (yz)\nplane. Therefore, for small values of the rotation, θ(I)\ndependence is linear. In principle, the parameter Γ can\nbe larger than one. In that case, when the condition\nI/Ic≥1/Γ is fulfilled, the magnetic moment will be ori-\nented along the y-axis. Therefore, applying a d.c. super-\nconducting current switches the direction of the magneti-\nzation, whereasapplyingana.c. currentona ϕ0-junction\ncould generate the precession of the magnetic moment.\nWe briefly comment on the situation when the direc-\ntion of the gradient of the spin-orbit potential is perpen-\ndicular (along y) to the easy axis z. To consider this\ncase we simply need to take ϕ0=ℓ(vso/vF)cosθ. The\ntotal energy (4) has two minima θ= (0,π), while apply-\ning the current removes the degeneracy between them.\nHowever, the energy barrier exists for the switch from\none minimum into another. This barrier may disappear\nif Γ>1 and the current is large enough I > Ic/Γ. In\nthis regime the superconducting current would provoke\nthe switching of the magnetization between one stable\nconfiguration θ= 0 and another θ=π. This corresponds\nto the transitions of the junction between + ϕ0and−ϕ0\nstates. The read-out of the state of the ϕ0-junction may\nbe easily performed if it is a part of some SQUID-like\ncircuit (the ϕ0-junction induces a shift of the diffraction\npattern by ϕ0).\nIn fact, the voltage-biased Josephson junction, and\nthus the a.c. Josephson effect provides an ideal tool\nto study magnetic dynamics in a ϕ0-junction. In such\na case, the superconducting phase varies with time like\nϕ(t) =ωJt[19]. IfℏωJ≪Tc, onecanusethestaticvalue\nfor the energy of the junction (4) considering ϕ(t) as an\nexternal potential. The magnetization dynamics are de-\nscribed by the Landau-Lifshitz-Gilbert equation (LLG)\n[20]\ndM\ndt=γM×Heff+α\nM0/parenleftbigg\nM×dM\ndt/parenrightbigg\n,(6)\nwhereHeff=−δF/VδMis the effective magnetic field\napplied to the compound, γthe gyromagnetic ratio, and\nαa phenomenological damping constant. The corre-\nsponding free energy F=Es+EMyields\nHeff=K\nM0/bracketleftbigg\nΓsin/parenleftbigg\nωJt−rMy\nM0/parenrightbigg\nˆy+Mz\nM0ˆz/bracketrightbigg\n,(7)\nwherer=ℓvso/vF. Introducing mi=Mi/M0,τ=ωFt3\n/s45/s49 /s48 /s49/s45/s49/s48/s49\n/s61 /s61/s49/s44 /s61/s48/s46/s49/s44 /s114/s61/s49/s109\n/s122\n/s109\n/s120/s99/s41\n/s45/s49 /s48 /s49/s45/s49/s48/s49\n/s61 /s61/s49/s44 /s61/s48/s46/s48/s53/s44 /s114/s61/s53/s109\n/s122\n/s109\n/s121 /s100/s41/s48 /s49 /s50 /s51/s45/s49/s48/s49/s109\n/s122/s40 /s41/s32/s110/s117/s109/s101/s114/s105/s99\n/s32/s97/s110/s97/s108/s121/s116/s105/s99\n/s61/s53 /s44 /s61/s53/s44 /s61 /s114/s61/s48/s46/s49/s97/s41\n/s45/s49 /s48 /s49/s45/s49/s48/s49\n/s61 /s61/s48/s46/s53/s44 /s61/s48/s46/s54/s44 /s114/s61/s49\n/s109\n/s121 /s109\n/s122/s98/s41/BDFIG. 2: Results of numerical analysis of the magnetic moment dynamics of the ϕ0-junction. a) Reversal of mzfrom analytical\nexpression Eq.(15) (dashed curve) and numerical one (norma l curve). The other curves are related to the Mtrajectory: b)\nin strong damping case c) and d) in the strong coupling regime revealing incomplete and complete magnetic moment reversa l,\nrespectively.\n(ωF=γK/M2\n0is the frequency of the ferromagnetic res-\nonance) in LLG equation (6) leads to\n\n\n˙mx=mz(τ)my(τ)−Γmz(τ)sin(ωτ−rmy)\n˙my=−mz(τ)mx(τ)\n˙mz= Γmx(τ)sin(ωτ−rmy),(8)\nwhereω=ωJ/ωF. The generalization of Eq.(8) for\nα/negationslash= 0 is straightforward. One considers first the ”weak\ncoupling” regime Γ ≪1 when the Josephson energy EJ\nis small in comparison with the magnetic energy EM. In\nthis case, the magnetic moment precess around the z-\naxis. If the other components verify ( mx,my)≪1, then\nthe equations (8) may be linearized, and the correspond-\ning solutions are\nmx(t) =ΓωcosωJt\n1−ω2andmy(t) =−ΓsinωJt\n1−ω2.(9)\nNear the resonance ωJ≈ωF, the conditions of lineariza-\ntion are violated and it is necessary to take the damping\nintoaccount.Theprecessingmagneticmomentinfluences\nthe current through the ϕ0-junction like\nI\nIc= sinωJt+Γr\n21\nω2−1sin2ωJt+...,(10)\ni.e., in addition to the first harmonic oscillations, the\ncurrent reveals higher harmonics contributions. The am-\nplitude oftheharmonicsincreasesneartheresonanceandchanges its sign when ωJ=ωF. Thus, monitoring the\nsecond harmonic oscillations of the current would reveal\nthe dynamics of the magnetic system.\nThe damping plays an important role in the dynamics\nof the considered system. It results in a d.c. contribu-\ntion to the Josephson current. Indeed, the corresponding\nexpressionfor my(t) in the presence of damping becomes\nmy(t) =ω+−ω−\nrsinωJt+α−−α+\nrcosωJt,(11)\nwhere\nω±=Γr\n2ω±1\nΩ±andα±=Γr\n2α\nΩ±,(12)\nwith Ω ±= (ω±1)2+α2. It thus exhibits a damped\nresonance as the Josephson frequency is tuned to the fer-\nromagnetic one ω→1. Moreover, the damping leads\nto the appearance of out of phase oscillations of my(t)\n(term proportional to cos ωJtin Eq.(11)). In the result\nthe current\nI(t)≈IcsinωJt+Icω+−ω−\n2sin2ωJt+\n+Icα−−α+\n2cos2ωJt+I0(α) (13)4\nacquires a d.c. component\nI0(α) =αΓr\n4/parenleftbigg1\nΩ−−1\nΩ+/parenrightbigg\n. (14)\nThis d.c. current in the presence of a constant voltage V\napplied to the junction means a dissipative regime which\ncan be easily detected. In some aspect, the peak of d.c.\ncurrent near the resonance is reminiscent of the Shapiro\nsteps effect in Josephson junctions under external r.f.\nfields. Note that the presence of the second harmonic\ninI(t) Eq.(13) should also lead to half-integer Shapiro\nsteps inϕ0-junctions [21].\nThe limit of the ”strong coupling” Γ ≫1 (butr≪1)\ncan also be treated analytically. In this case, my≈0 and\nsolutions of Eq.(8) yields\n\n\nmx(t) = sin/bracketleftbiggΓ\nω(1−cosωJt)/bracketrightbigg\nmz(t) = cos/bracketleftbiggΓ\nω(1−cosωJt)/bracketrightbigg,(15)\nwhich are the equations of the magnetization reversal, a\ncomplete reversal being induced by Γ /ω > π/2. Strictly\nspeaking, these solutions are not exact oscillatory func-\ntions in the sense that mz(t) turns around the sphere\ncenter counterclockwise before reversingits rotation, and\nreturns to the position mz(t= 0) = 1 clockwise, like a\npendulum in a spherical potential (see Fig.2.c).\nFinally, we have performed numerical studies of the\nnon-linear LLG Eq.(6) for some choices of the parame-\nters when the analytical approaches fail. To check the\nconsistency of our numerical and analytical approaches,\nwe present in Fig.2.a) the corresponding mz(t) depen-\ndences for low-damping regimes. They clearly demon-\nstrate the possibility of the magnetization reversal. In\nFigs.2b-d), some trajectories of the magnetization vec-\ntorsarepresentedforgeneralcouplingregimes. Thesere-\nsults demonstrate that the magnetic dynamics of S/F/S\nϕ0-junctionmaybe prettycomplicatedand stronglynon-\nharmonic.\nIf theϕ0-junction is exposed to a microwave radia-\ntion at angular frequency ω1, the physics that emerge\nare very rich. First, in addition to the Shapiro steps\natωJ=nω1, half-integer-steps will appear. Secondly,\nthe microwave magnetic field may also generate an addi-\ntional magnetic precession with ω1frequency. Depend-\ning on the parameters of ϕ0-junction and the amplitude\nof the microwave radiation the main precession mecha-\nnism may be related either to the Josephson current or\nthe microwave radiation. In the last case the magnetic\nspin-orbit coupling may substantially contribute to the\namplitude of the Shapiro steps. Therefore, we could ex-\npect a dramatic increase of this amplitude at frequencies\nnear the ferromagnetic resonance. When the influence\nof the microwave radiation and Josephson current on theprecession is comparable, a very complicated regime may\nbe observed.\nIn the present work we considered the case of the easy-\naxis magnetic anisotropy. If the ferromagnet presents an\neasy-planeanisotropythanqualitativelythemainconclu-\nsions ofthis article remain the same because the coupling\nbetween magnetism and superconductivity depends only\non theMycomponent. However, the detailed dynamics\nwould be strongly affected by a weak in-plane anisotropy.\nTo summarize, we have demonstrated that S/F/S ϕ0-\njunctions provide the possibility to generate magnetic\nmoment precession via Josephson current. In the regime\nof strong coupling between magnetization and current,\nmagnetic reversal may also occur. These effects have\nbeen studied analytically and numerically. We believe\nthat the discussed properties of the ϕ0-junctions could\nopen interesting perspectives for its applications in spin-\ntronics.\nThe authors are grateful to Z. Nussinov, J. Cayssol,\nM. Houzet, D. Gusakova, M. Roche and D. Braithwaite\nfor useful discussions and comments. This work was sup-\nported by the French ANR Grant N◦ANR-07-NANO-\n011: ELEC-EPR.\n∗Electronic address: f.konschelle@cpmoh.u-bordeaux1.fr\n†Also atInstitut Universitaire de France .\n[1] I.ˇZuti´ c, J. Fabian, and S. Das Sarma. Rev. Mod. Phys.\n76, 323 (2004).\n[2] J. Hauptmann, J. Paaske, and P. Lindelof. Nature\nPhysics4, 373 (2008).\n[3] R. Winkler. Spin-orbit coupling effects in two-\ndimensional electron and hole systems (Springer, New\nYork, 2003).\n[4] A. Buzdin. Phys. Rev. Lett. 101, 107005 (2008).\n[5] A.A.Reynoso et al.Phys.Rev.Lett. 101, 107001 (2008).\n[6] A. I. Buzdin. Rev. Mod. Phys. 77, 935 (2005).\n[7] C. Bell et al.Phys. Rev. Lett. 100, 047002 (2008).\n[8] J.-X. Zhu, and A. V. Balatsky. Phys. Rev. B 67, 174505\n(2003).\n[9] L. Bulaevskii et al.Phys. Rev. Lett. 92, 177001 (2004).\n[10] J.-X. Zhu et al.Phys. Rev. Lett. 92, 107001 (2004).\n[11] Z. Nussinov et al.Phys. Rev. B 71, 214520 (2005).\n[12] S. Takahashi et al.Phys. Rev. Lett. 99, 057003 (2007).\n[13] M. Houzet. Phys. Rev. Lett. 101, 057009 (2008).\n[14] X. Waintal, and P. W. Brouwer. Phys. Rev. B 65, 054407\n(2002).\n[15] V. Braude, and Y. M. Blanter. Phys. Rev. Lett. 100,\n207001 (2008).\n[16] K. K. Likharev. Dynamics of Josephson junctions and\ncircuits(Gordon and Beach Science Publishers, 1986).\n[17] S. Hikino et al.J. Phys. Soc. Jap. 77, 053707 (2008).\n[18] A. Y. Rusanov et al.Phys. Rev. Lett. 93, 057002 (2004).\n[19] B. D. Josephson. Superconductivity (in two volumes),\nvol. 1, chapter 9, Weakly coupled superconductors (R.D\nParks, Marcel Dekker, Inc., 1968).5\n[20] E. M. Lifshitz, and L. P. Pitaevskii. Course of theoret-\nical physics, Volume 9 : Theory of the condensed state\n(Butterworth Heinemann, 1991).[21] H. Sellier et al.Phys. Rev. Lett. 92, 257005 (2004)." }, { "title": "1905.09241v1.Floquet_Second_Order_Topological_Superconductor_Driven_via_Ferromagnetic_Resonance.pdf", "content": "Floquet Second-Order Topological Superconductor\nDriven via Ferromagnetic Resonance\nKirill Plekhanov, Manisha Thakurathi, Daniel Loss, and Jelena Klinovaja1\n1Department of Physics, University of Basel, Klingelbergstrasse 82, CH-4056 Basel, Switzerland\n(Dated: May 23, 2019)\nWe consider a Floquet triple-layer setup composed of a two-dimensional electron gas with spin-\norbit interactions, proximity coupled to an s-wave superconductor and to a ferromagnet driven\nat resonance. The ferromagnetic layer generates a time-oscillating Zeeman \feld which competes\nwith the induced superconducting gap and leads to a topological phase transition. The resulting\nFloquet states support a second-order topological superconducting phase with a pair of localized\nzero-energy Floquet Majorana corner states. Moreover, the phase diagram comprises a Floquet\nhelical topological superconductor, hosting a Kramers pair of Majorana edge modes protected by an\ne\u000bective time-reversal symmetry, as well as a gapless Floquet Weyl phase. The topological phases\nare stable against disorder and parameter variations and are within experimental reach.\nIntroduction. Over the last decade topological states\nof matter [1{5] have attracted a lot of attention. Re-\ncently, particular interest has been raised by higher-\norder topological insulators and superconductors [6{8],\nwhich host topologically protected gapless modes on their\nhigher-order faces ( e.g. corners in d > 1, hinges in\nd > 2, withdbeing the spatial dimension of the sys-\ntem). However, such systems were studied mostly at the\nstatic level [9{19], with the out-of-equilibrium dynam-\nics [20{26] rarely addressed. At the same time, Floquet\nengineering [1{4], based on applying time-periodic per-\nturbations, has been serving as a powerful tool to gener-\nate exotic phases of quantum matter, including topolog-\nical and Chern insulators [31{50], as well as non-Abelian\nstates such as Majorana bound states (MBSs) [51{59].\nSuch Floquet phases can be generated by time-dependent\nelectromagnetic \felds [31{38, 57, 58]. Strong oscillating\nelectric \felds are easy to obtain for this purpose, but\nthe direct coupling to the spin degree of freedom is more\nchallenging, typically achieved only indirectly via strong\nspin orbit interaction [60, 61]. Hence, \fnding ways to\ngenerate strong oscillating magnetic \felds which couple\ndirectly to the spin is important. One possible solution\nconsists in using magnetic proximity e\u000bects, which, for\nthe static case, have already been widely studied for su-\nperconductors [62, 63] and topological insulators [64{66].\nIn this work we consider a driven triple-layer setup\nwhich allows us to engineer a Floquet higher-order topo-\nlogical superconductor (FHOTS). Its key component is\na two-dimensional electron gas (2DEG) with spin-orbit\ninteraction (SOI) sandwiched between an s-wave super-\nconductor (SC) and a ferromagnet (FM), as shown in\nFig. 1(a). The FM layer is resonantly driven by an ex-\nternal \feldH(t), which results in the generation of an\noscillating magnetic \feld B(t), giving rise to strong Zee-\nman coupling in the 2DEG. The out-of-plane component\nofB(t) competes with the proximity-induced supercon-\nducting gap and leads to a topological phase transition\nin the 2DEG. The topological phase is identi\fed as Flo-\nquet helical topological superconductor (FHeTS), char-\nacterized by the presence of a Kramers pair of gapless\nFIG. 1. (a) Schematics of the Floquet triple-layer setup con-\nsisting of a 2DEG (blue layer) with SOI of strengths \u000bx\nand\u000by, proximity coupled to an s-wave SC (red layer) at\nthe bottom and a FM at the top (green layer) which is\nresonantly driven by an external time-dependent magnetic\n\feldH(t) to generate an oscillating magnetic \feld B(t) =\nB?cos(!t)ez+Bksin(!t)ukin the 2DEG. (b) Band structure\nof the 2DEG in the isotropic regime \u000bx=\u000by. The chemical\npotential\u0016is \fxed below the crossing point of the spin-split\nbands (indicated by the blue area). The driving frequency !\nis tuned to achieve resonance at the smallest Fermi momen-\ntum (represented by the two green circles) between the two\nFloquet bands labeled by \u001c=\u00061.\nhelical Majorana edge modes protected by an e\u000bective\ntime-reversal symmetry. A moderate in-plane component\nofB(t) opens a gap in the helical edge modes through\na mass term that changes sign at the corners in a rect-\nangular geometry, resulting in Majorana corner states\n(MCSs) { a distinct feature of the FHOTS. In the regime\nwhere the in-plane component of B(t) is strong, the sys-\ntem moves into a gapless Weyl phase. The topological\nphase diagram is robust against moderate local disorder,\ndetuning from resonance, and static magnetic \felds.\nModel. The triple-layer setup is composed of a 2DEG\nwith strong Rashba and Dresselhaus SOIs, proximity\ncoupled to an s-wave SC and a FM, as shown in Fig. 1(a).\nWe assume that the SOI vector is along zdirection, which\nis perpendicular to the xy2DEG-plane. The SOI cou-\npling strengths \u000bx=\u000bR+\fDand\u000by=\u000bR\u0000\fDare\nexpressed in terms of the Rashba and Dresselhaus SOI\ncoe\u000ecients \u000bRand\fDfor a proper choice of the coordi-\nnate system [67{74]. Introducing a creation operator y\n\u001bkarXiv:1905.09241v1 [cond-mat.mes-hall] 22 May 20192\nacting on an electron with momentum k= (kx;ky) and\nspin component \u001balong thezaxis, the corresponding\nHamiltonian reads\nH0=X\n\u001b\u001b0Z\ndk y\n\u001bk\u0014~2k2\n2m\u0000\u0016 (1)\n\u0000\u000bxkx\u001by+\u000byky\u001bx\u0015\n\u001b\u001b0 \u001b0k:\nHere,\u001bjare the Pauli matrices acting in spin space.\nThe chemical potential \u0016is calculated from the cross-\ning point of the two spin-split bands at k= 0. In\nthe following, it will be convenient to introduce the\nSOI energy Eso=~2k2\nso=(2m) and the SOI momentum\nkso=m\u000bx=~2. We note that this Hamiltonian e\u000bectively\ndescribes a 2D array of coupled Rashba wires if the mass\nmis also chosen to be anistropic in the xyplane such\nthatmy6=mx[38, 75{86].\nThe proximity e\u000bect between the 2DEG and the SC is\ndescribed by the following Hamiltonian\nHsc=\u0001sc\n2X\n\u001b\u001b0Z\ndk\u0010\n y\n\u001bk[i\u001by]\u001b\u001b0 y\n\u001b0(\u0000k)+ H.c.\u0011\n;(2)\nwhere \u0001 scis the induced SC gap. The resulting 2DEG-\nSC heterostructure is placed in the vicinity of a FM layer,\nand the setup is subjected to an external magnetic \feld\nH(t). Under the FM resonance condition (see discussion\nbelow), the FM generates an oscillating demagnetizing\n\feld which adds up to H(t) to produce a total magnetic\n\feldB(t) =B?cos(!t)ez+Bksin(!t)ukin the 2DEG.\nHere,!= 2\u0019=T denotes the oscillation frequency and\nthe 2D vector uk= (ux;uy) indicates the orientation of\nthe magnetic \feld in the xyplane. The FM proximity\ne\u000bect is described by the following Floquet-Zeeman term\nHZ(t) = 2X\n\u001b\u001b0Z\ndk y\n\u001bk\u0010\nt?\nZcos(!t)[\u001bz]\u001b\u001b0\n+tk\nZsin(!t)[uk\u0001\u001b]\u001b\u001b0\u0011\n \u001b0k;(3)\nwheret\u0017\nZ=\u0016Bg\u0017B\u0017=2 (with\u0017=k;?) are two Floquet-\nZeeman amplitudes. The anisotropy in the g-factors,\nwhich leads to gkandg?, arises from the quantum con-\n\fnement and the intrinsic strain of the 2DEG [87{89].\nThe resulting time-dependent problem can be solved\nusing the Floquet formalism [1{4], by writing the quasi-\nenergy operator H=H0+Hsc+HZ(t)\u0000i~@tin the\nFloquet-Hilbert space generated by T-periodic states\n l\u001bk= exp(\u0000il!t) \u001bk,l2Z. In this basis Hacquires\na simple block-diagonal form, where each block, also re-\nferred to as a Floquet band, is composed of the modes\nwith the same index l. The static term acts within the\nsame block and receives an additional constant energy\nshift~!l, while the oscillating term couples di\u000berent\nblocks. We assume that the chemical potential is re-\nstricted to\u0000Eso< \u0016 < 0, so that the frequency !can\nbe resonantly tuned to the energy di\u000berence between the\n(a) (b)\ntrivialFHeTS Weyl\nFIG. 2. (a) Energy spectrum E(kx) in the FHeTS phase\nwitht?\nZ=\u0001sc= 3 and\u000bx=\u000by. The helical edge modes are\nlocalized at the edges. (b) Phase diagram showing the gap to\nthe \frst excited bulk state (in units of Eso) as a function of\nthe ratios\u000by=\u000bxandt?\nZ=\u0001sc. Blue lines indicate the phase\nboundaries. In the isotropic regime, \u000bx=\u000by, the phase tran-\nsition occurs at the critical point \u0001 sc=t?\nZ. When\u000bx6=\u000by,\nthe critical point is transformed into a gapless Weyl phase,\nand the value t?\nZ=\u0001screquired to reach the FHeTS increases\nup to the point of strong anisotropy beyond which the FHeTS\ncannot be reached. Remaining parameters in both simula-\ntions aretk\nZ= 0, \u0001 sc=Eso= 0:2,ksoLx=ksoLy= 80, and\n\u0016=\u0000Eso=2.\ntwo spin-split bands. This allows us to treat the oscil-\nlating terms at low energies by only taking into account\nthe coupling between the modes at l= 0, to which we\nassociate a Floquet band index \u001c= 1, and the modes at\nl=\u00001 with\u001c=\u00161. As a result, in the Nambu basis \ty\nk\n= ( y\n1\"k, y\n1#k, y\n\u00161\"k, y\n\u00161#k, 1\"\u0000k, 1#\u0000k, \u00161\"\u0000k, \u00161#\u0000k),\nthe total Hamiltonian reads as H=R\ndk\ty\nkHk\tk=2,\nwith the Hamiltonian density\nHk=\u0014~2k2\n2m\u0000\u0016+~!(\u001cz\u0000\u001c0)\n2\u0015\n\u0011z+ \u0001 sc\u0011y\u001by+\u000byky\u001bx\n\u0000\u000bxkx\u0011z\u001by+t?\nZ\u0011z\u001cx\u001bz+tk\nZ(ux\u001cy\u001bx+uy\u0011z\u001cy\u001by) (4)\nwith the Pauli matrices \u001ci(\u0011i) acting in the Floquet\n(particle-hole) space. In the following, we analyze dif-\nferent topological phases of the system as a function of\nthe parameters appearing in Eq. (4).\nFloquet Helical Topological Superconductor. In order\nto determine the phase diagram of our model, we \frst\nconsider the e\u000bect of the out-of-plane component of B(t)\nby imposing tk\nZ= 0 in Eq. (4). In the isotropic regime\nwith\u000bx=\u000by, the Fermi surface is composed of two con-\ncentric circles and the problem depends only on the mag-\nnitude of the momentum jkj, as shown in Fig. 1(b). The\nresonance condition for the frequency !is satis\fed along\nthe entire circle with the smallest Fermi momentum (see\nthe Supplemental Material (SM) [90] for more details).\nThe Hamiltonian is linearized close to the Fermi surface\nand provides the eigenenergies E2\n1= (~vF\u000ek)2+ \u00012\nscand\nE2\n2;\u0006= (~vF\u000ek)2+(t?\nZ\u0006\u0001sc)2, withvF=\u000bx=~the Fermi\nvelocity and \u000ekthe radial distance from the Fermi sur-\nface. The phase diagram consists of two gapped phases\nseparated by the gapless line \u0001 sc=t?\nZ. The topologically3\ntrivial (topological) phase is identi\fed with the regime\n\u0001sc> t?\nZ(\u0001sc< t?\nZ). Fortk\nZ= 0, the system is char-\nacterized by an emerging e\u000bective time-reversal symme-\ntryTe\u000b=\u0000i\u001cz\u001byK, a particle-hole symmetry P=\u0011xK,\nand a chiral symmetry UC=PTe\u000b, withKthe complex-\nconjugation operator. Thus, the system belongs to the\nDIII symmetry class with Z2topological invariant.\nThe topological phase, denoted as FHeTS, hosts gap-\nless boundary modes { a Kramers pair of Floquet Ma-\njorana fermions j\b\u0006i, obeying Pj\b\u0006i=j\b\u0006iand\nTe\u000bj\b\u0006i=\u0006j\b\u0007i. The Kramers partners propagate in\nopposite directions along the same edge, forming a pair\nof helical modes protected by the e\u000bective time-reversal\nand particle-hole symmetries. We have veri\fed the pres-\nence of these modes numerically in the discretized version\nof the model (see the SM [90]) de\fned on a rectangular\nlattice of size Lx\u0002Lywith periodic boundary conditions\nalongx, as shown in Fig. 2(a).\nIf the rotation symmetry is broken ( \u000bx6=\u000by), the res-\nonance condition can be satis\fed only along a particular\ndirection in momentum space, resulting in an o\u000b-set \u000e!\nin the resonance condition almost everywhere except at\na few points on the Fermi surface. While a small \u000e!in\nthe weak anisotropy regime hardly a\u000bects the phase di-\nagram and can be compensated by increasing t?\nZ, strong\nanisotropy e\u000bects are more drastic. In Fig. 2(b), we cal-\nculate numerically the energy of the lowest bulk state as\na function of the ratios \u000by=\u000bxandt?\nZ=\u0001sc. We see that\nthe critical point \u0001 sc=t?\nZtransforms into a gapless Weyl\nphase at a \fnite value of anisotropy ( \u000by\u0000\u000bx)=\u000bx. This\ngapless regime is characterized by a semi-metal energy\nstructure with four Weyl cones. The nodes of the Weyl\ncones appear \frst on the Fermi surface and move further\nin the reciprocal space, when the parameters t?\nZ=\u0001scand\n\u000by=\u000bxare modi\fed. We also note that if we decrease the\nratiot?\nZ=\u0001scup to a point of reaching the phase transi-\ntion to the trivial phase, the low energy physics becomes\ninsensitive to the anisotropy.\nFloquet Majorana corner states. Next, we analyze the\ne\u000bect of an oscillating in-plane magnetic \feld that breaks\nthe e\u000bective time-reversal symmetry Te\u000band, thus, gaps\nout the helical edge modes of the FHeTS. Nevertheless,\nthe system remains topologically non-trivial as it now\nhosts a set of zero-energy MCSs, characteristic for the\nFHOTS phase [91]. The presence of such MCSs is uncov-\nered by focussing on the low-energy degrees of freedom\nexpressed in terms of the Majorana edge modes j\b\u0006i.\nThe in-plane Zeeman \feld Bkcouples the two helical\nmodes, leading to the following low-energy Hamiltonian\ndensity:\nHedge=~ve\nFjkj\u001az+~tk\nZ\u001ay; (5)\nwhere the Pauli matrices \u001aiact in the space of\nj\b\u0006i,ve\nFis the velocity of the helical edge modes,\nand~tk\nZ=tk\nZImh\b\u0000jux\u001cy\u001bx+uy\u0011z\u001cy\u001byj\b+iis the `mass\n(a)\n(d)\nWeyl\ntrivialFHOTSFHeTS\n(c)\n(b)\nFIG. 3. (a,b,d) Probability density of the lowest energy state\nin the FHOTS phase for t?\nZ=\u0001sc= 3,tk\nZ=\u0001sc= 2:5, and\n\u000bx=\u000by. The inset shows the 10 lowest eigenenergies. (a)\nForuk= (1;1)=p\n2, the in-plane Zeeman \feld Bkopens gaps\nin all edge modes such that zero-energy MCSs emerge at two\nopposite corners. (b) The vector uk= (0;1) is parallel to the\nedges along which the system stays gapless. (c) Phase dia-\ngram showing the bulk gap (color coded in units of Eso) as a\nfunction of the ratios t?\nZ=\u0001scandtk\nZ=\u0001scwithuk= (1;1)=p\n2.\nThe critical point \u0001 sc=t?\nZattk\nZ= 0 merges into a gapless\nWeyl phase at \fnite tk\nZ. As a result, higher value of t?\nZ=\u0001sc\nare required to reach the topological phase. The various\nphase boundaries correspond to \u000by=\u000bx= 0:8 (red dotted),\n0:9 (purple dashed), and 1 :0 (blue solid). (d) The MCSs are\nstable against moderate external perturbations and disorder:\n\u000e!=p\n3S\u0016= \u0001 Z= 0:10Eso. Rest of parameters are the\nsame as in Fig. 2.\nterm'. Thus, our system is described by the well-known\nJackiw-Rebbi model [7, 8].\nFrom the symmetry of the modes j\b\u0006i, we deduce that\nthe value of the mass term ~tk\nZonly depends on the com-\nponent of the in-plane \feld Bkwhich is perpendicular to\nthe corresponding edge in a rectangular geometry (see\nthe SM [90]). Generally, the sign of ~tk\nZis opposite on two\nparallel edges at x= 0 (y= 0) andx=Lx(y=Ly).\nHence, ~tk\nZhas to change its sign at two opposite corners\nof the 2DEG, leading to the emergence of domain walls at\nthese corners that host zero-energy MCSs, see Fig. 3(a).\nIn the special case when ukis parallel to one of the edges,\nthe corresponding edge modes stay gapless, see Fig. 3(b).\nThe simple boundary description in terms of the\nJackiw-Rebbi model is expected to work in the regime\nwhere the amplitude of the in-plane magnetic \feld is\nsmall. In order to construct the full phase diagram, we\ncalculated numerically the gap to the \frst excited bulk\nstate as a function of the ratios t?\nZ=\u0001scandtk\nZ=\u0001sc, see\nFig. 3(c). The FHOTS phase emerges from the FHeTS\nphase at non-zero tk\nZ. However, if tk\nZis large, the system\nenters into the gapless Weyl phase. To observe MCSs, the4\nFM\n2DEG\nSC\nFIG. 4. Setup of the FM layer to generate the desired\noscillating magnetic \feld in the 2DEG. The 2DEG-SC het-\nerostructure is placed in the vicinity of the FM. The system is\nsubjected to an external magnetic \feld H(t) =H0+h(t) (yel-\nlow lines) generating FM resonance (shown not to scale). This\ninduces the precession of the FM magnetization M(t) (black\nlines inside the FM) and demagnetizing \feld D(t) (black lines\noutside of the FM). Close to the surface of the FM layer, the\nstatic components of D(t) and H(t) cancel out, so that the\ndynamics of the total \feld B(t) =H(t) +D(t) in the 2DEG\nlayer is close to a full 360\u000erotation.\nFloquet-Zeeman \feld perpendicular to the 2DEG plane\nshould dominate such that the condition tk\nZ< t?\nZis ful-\n\flled. All the results remain qualitatively the same for a\nweak anisotropy ( \u000bx6=\u000by), which shifts the topological\nphase transition line only slightly.\nExperimental feasibility. Next, we discuss the stabil-\nity of our setup. We check numerically that the topo-\nlogical phases are robust against an o\u000b-set \u000e!in the res-\nonance frequency !. Similarly, we check the stability\nwith respect to an on-site disorder by adding a \ructu-\nating chemical potential randomly chosen from a uni-\nform distribution with standard deviation S\u0016and with\nrespect to a static magnetic \feld by adding a Zeeman\nterm with strength \u0001 Zdirected both in-plane and out-of-\nplane. The result of the calculations is shown in Fig. 3(d).\nThe topological phases are stable against the perturba-\ntions of a strength comparable to the gap. The e\u000bect of\nthe out-of-plane component of the static magnetic \feld\nis also less important: the MCSs can be observed even\nfor \u0001 Z\u0018t?\nZand disappear only when the static term\nbecomes stronger.\nIn experiments, the proximity induced SC gap \u0001 scis\nexpected to be of the order of 0 :05 meV, depending on\nthe properties of the SC and the strength of the cou-\npling between the SC and the 2DEG [94{96]. As shown\nin this work, the strength of the Floquet-Zeeman am-\nplitudet?\nZshould exceed \u0001 scto reach the topological\nregimes. Hence, assuming that the 2DEG material has\nan electron g-factorg?= 15, the amplitude of the mag-\nnetic \feldB?should be of the order of 0 :1 T. At the\nsame time, the static component of the magnetic \feld\nshould be smaller than the dynamic one and the oscilla-\ntion frequency !should be in the GHz range. The FMlayer in the setup is proposed to generate the required\nZeeman \felds as follows. Applying an external magnetic\n\feldH(t) =H0+h(t) withjh(t)j\u001cjH0jandh(t)?H0\nunder FM resonance condition induces a precession of the\nFM magnetization M(t) [11, 12]. The precession cone of\nM(t) depends on the angle between the FM easy axis and\nH0, while the resonance frequency is determined by the\nmagnitudejH0j. Outside of the FM, the total \feld B(t)\nis equal to the sum of the external \feld H(t) and an os-\ncillating demagnetizing \feld D(t), see Fig. 4. Hence, by\ncarefully choosing the system geometry, the static com-\nponent ofB(t) could be adjusted close to zero over a\nlarge region of space in the proximity of the FM surface\nincluding the 2DEG. The amplitude of the remaining os-\ncillating component overcomes the threshold of 0 :1 T in-\nside the 2DEG, as we have con\frmed by micromagnetic\nsimulations (see the SM [90]). Promising candidates for\nsuch FMs are e.g. EuS [64], GdN [65], and YiG [66].\nAlternatively, the fast switching or the sustained oscil-\nlation of the FM magnetization has already been achieved\nexperimentally by shining optical light on a FM (via an\nall-optical magnetization reversal) [99{101], by applying\npiezostrain [102, 103] or by injecting a spin-polarized cur-\nrent (via a spin-orbit torque) [104{108]. This domain of\nresearch is currently under an active exploration because\nof its crucial role in the implementation of magnetic mem-\nory and logic devices. We also note that in our setup, the\nmagnetic \feld in the 2DEG can originate from both the\nFM demagnetizing \feld at a moderate range and from\nthe exchange interactions at atomic distances.\nConclusions. We have considered a Floquet triple-\nlayer setup of a 2DEG proximity coupled to a SC and\na FM. Under resonant drive the FM induces an oscillat-\ning Zeeman \feld in the 2DEG. The out-of-plane com-\nponent of the magnetic \feld competes with the proxim-\nity induced SC gap and leads to the emergence of the\nFHeTS hosting an e\u000bective Kramers pair of gapless he-\nlical edge modes. Moreover, the in-plane component of\nthe magnetic \feld enters into the low-energy description\ncorresponding to the e\u000bective Jackiw-Rebbi model as a\nmass term and opens a gap in the edge mode spectrum.\nChange in the sign of the mass term, which inevitably\noccurs at two opposite corners of the system in a rectan-\ngular geometry, leads to the emergence of Floquet MCSs.\nWe argued that the proposed setup is within experimen-\ntal reach combining available magnetic, semiconducting,\nand superconducting materials.\nAcknowledgments. We acknowledge very much the\ndiscussions with Patrick Maletinsky, Martino Poggio,\nChristina Psaroudaki, Marko Ran\u0014 ci\u0013 c, and Flavio Ronetti.\nThis work was supported by the Swiss National Science\nFoundation, NCCR QSIT, and the Georg H. Endress\nfoundation. This project received funding from the Euro-\npean Union's Horizon 2020 research and innovation pro-\ngram (ERC Starting Grant, grant agreement No 757725).5\n1M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045\n(2010).\n2X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057\n(2011).\n3M. Sato and Y. Ando, Rep. Prog. Phys. 80, 076501 (2017).\n4J. Wang and S.-C. Zhang, Nat. Mat. 16, 1062-1067 (2017).\n5X.-G. Wen, Rev. Mod. Phys. 89, 41004 (2017).\n6W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Sci-\nence 357, 6346 (2017).\n7W. A. Benalcazar, B. A. Bernevig, and T. L. Hughes, Phys.\nRev. B 96, 245115 (2017).\n8Z. Song, Z. Fang, and C. Fang, Phys. Rev. Lett. 119,\n246402 (2017).\n9Y. Peng, Y. Bao, and F. von Oppen, Phys. Rev. B 95,\n235143 (2017).\n10S. Imhof, C. Berger, F. Bayer, J. Brehm, L. W.\nMolenkamp, T. Kiessling, F. Schindler, C. Hua Lee, M.\nGreiter, T. Neupert, and R. Thomale, Nat. Phys. 14, 925\n(2018).\n11M. Geier, L. Trifunovic, M. Hoskam, and P. W. Brouwer,\nPhys. Rev. B 97, 205135 (2018).\n12F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S.\nP. Parkin, B. A. Bernevig, and T. Neupert, Science Adv.\n4, 6 (2018).\n13C.-H. Hsu, P. Stano, J. Klinovaja, and D. Loss, Phys. Rev.\nLett. 121, 196801 (2018).\n14T. Liu, J. J. He, and F. Nori, Phys. Rev. B 98, 245413\n(2018).\n15Y. Volpez, D. Loss, and J. Klinovaja, Phys. Rev. Lett. 122,\n126402 (2019).\n16M. Ezawa, Scienti\fc Reports 9, 5286 (2019).\n17A. Agarwala, V. Juricic, B. Roy, arXiv:1902.00507.\n18K. Laubscher, D. Loss, and J. Klinovaja, arXiv:1905.00885.\n19D. Calugaru, V. Juricic, and B. Roy, Phys. Rev. B 99,\n041301(R) (2019).\n20S. Franca, J. van den Brink, and I. C. Fulga, Phys. Rev. B\n98, 201114(R) (2018).\n21R. W. Bomantara, L. Zhou, J. Pan, and J. Gong, Phys.\nRev. B 99, 045441 (2019).\n22Y. Peng and G. Refael, arXiv:1811.11752 (2018).\n23B. Huang and W. V. Liu, arXiv:1811.00555 (2018).\n24M. Rodriguez-Vega, A. Kumar, and B. Seradjeh,\narXiv:1811.04808 (2018).\n25R. Seshadri, A. Dutta, and D. Sen, arXiv:1901.10495\n(2019).\n26T. Nag, V. Juricic, B. Roy, arXiv:1904.07247 (2019).\n27J. H. Shirley, Phys. Rev. 138, 979 (1965).\n28N. Goldman and J. Dalibard, Phys. Rev. X 4, 031027\n(2014).\n29M. Bukov, L. D'Alessio, and A. Polkovnikov, Advances in\nPhysics 64, 2, 139 (2015).\n30A. Eckardt and E. Anisimovas, New. J. Phys. 17, 093039\n(2015).\n31T. Oka and H. Aoki, Phys. Rev. B 79, 081406(R) (2009).\n32J.-I. Inoue and A. Tanaka, Phys. Rev. Lett. 105, 017401\n(2010).\n33T. Kitagawa, T. Oka, A. Brataas, L. Fu, and E. Demler,\nPhys. Rev. B 84, 235108 (2011).\n34N. H. Lindner, G. Refael, and V. Galitski, Nat. Phys. 7,\n490 (2011).35N. H. Lindner, D. L. Bergman, G. Refael, and V. Galitski,\nPhys. Rev. B 87, 235131 (2013).\n36J. Cayssol, B. D\u0013 ora, F. Simon, and R. Moessner, Phys.\nStatus Solidi RRL, 7, 101 (2013).\n37P. M. Perez-Piskunow, G. Usaj, C. A. Balseiro, and L. E.\nF. Foa Torres, Phys. Rev. B 89, 121401(R).\n38J. Klinovaja, P. Stano, and D. Loss, Phys. Rev. Lett. 116,\n176401 (2016).\n39T. Kitagawa, M. A. Broome, A. Fedrizzi, M. S. Rudner,\nE. Berg, I. Kassal, A. Aspuru-Guzik, E. Demler, and A.\nG. White, Nat. Comm. 3, 882 (2012).\n40M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D.\nPodolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit,\nNature 496, 196 (2013).\n41M. Aidelsburger, M. Atala, M. Lohse, J. T. Barreiro,\nB. Paredes, and I. Bloch, Phys. Rev. Lett. 111, 185301\n(2013).\n42G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat, T.\nUehlinger, D. Greif, and T. Esslinger, Nature 515, 237\n(2014).\n43K. Le Hur, L. Henriet, A. Petrescu, K. Plekhanov, G.\nRoux, and M. Schir\u0013 o, C. R. Physique 17, 808 (2016).\n44T. Kitagawa, E. Berg, M. Rudner, and E. Demler, Phys.\nRev. B 82, 235114 (2010).\n45M. S. Rudner, N. H. Lindner, E. Berg, and M. Levin, Phys.\nRev. X 3, 031005 (2013).\n46R. Roy and F. Harper, Phys. Rev. B 96, 155118 (2017).\n47L. J. Maczewsky, J. M. Zeuner, S. Nolte, and A. Szameit,\nNat. Comm. 8, 13756 (2017).\n48S. Yao, Z. Yan, and Z. Wang, Phys. Rev. B 96, 195303\n(2017).\n49G. M. Graf and C. Tauber, Annales Henri Poincar\u0013 e, 19,\n709 (2018).\n50A. G. Grushin, \u0013A. G\u0013 omez-Le\u0013 on, and T. Neupert, Phys.\nRev. Lett. 112, 156801 (2014).\n51L. Jiang, T. Kitagawa, J. Alicea, A. R. Akhmerov, D.\nPekker, G. Refael, J. I. Cirac, E. Demler, M. D. Lukin,\nand P. Zoller, Phys. Rev. Lett. 106, 220402 (2011).\n52A. A. Reynoso and D. Frustaglia, Phys. Rev. B 87, 115420\n(2013).\n53M. Thakurathi, A. A. Patel, D. Sen, and A. Dutta, Phys.\nRev. B 88, 155133 (2013).\n54D. E. Liu, A. Levchenko, and H. U. Baranger, Phys. Rev.\nLett. 111, 047002 (2013).\n55A. Kundu and B. Seradjeh, Phys. Rev. Lett. 111, 136402\n(2013).\n56M. Thakurathi, K. Sengupta, and D. Sen, Phys. Rev. B\n89, 235434 (2014).\n57M. Thakurathi, D. Loss, and J. Klinovaja, Phys. Rev. B\n95, 155407 (2017).\n58D. M. Kennes, N. M uller, M. Pletyukhov, C. Weber, C.\nBruder, F. Hassler, J. Klinovaja, D. Loss, H. Schoeller,\narXiv:1811.12062 (2018).\n59D. T. Liu, J. Shabani, and A. Mitra, Phys. Rev. B 99,\n094303 (2019).\n60E. I. Rashba, Sov. Phys. Solid State 2, 1109 (1960).\n61C. Kloe\u000bel, D. Loss, Annu. Rev. Condens. Matter Phys.\n4, 51 (2013).\n62A. I. Buzdin, Rev. Mod. Phys. 77, 935 (2005).\n63B. Li, N. Roschewsky, B. A. Assaf, M. Eich, M. Epstein-\nMartin, D. Heiman, M. M unzenberg, and J. S. Moodera,6\nPhys. Rev. Lett. 110, 097001 (2013).\n64F. Katmis, V. Lauter, F. S. Nogueira, B. A. Assaf, M. E.\nJamer, P. Wei, B. Satpati, J. W. F., I. Eremin, D. Heiman,\nP. Jarillo-Herrero, and J. S. Moodera, Nature 533, 513\n(2016).\n65A. Kandala, A. Richardella, D. W. Rench, D. M. Zhang,\nT. C. Flanagan, and N. Samarth, Appl. Phys. Lett. 103,\n202409 (2013).\n66Z. Jiang, C.-Z. Chang, C. Tang, J.-G. Zheng, J. S. Mood-\nera, and J. Shi, AIP Advances 6, 055809 (2016).\n67J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys.\nRev. Lett. 78, 1335 (1997).\n68G. Engels, J. Lange, T. Schapers, and H. Luth, Phys. Rev.\nB55, R1958 (1997).\n69J. Schliemann, J. C. Egues, and D. Loss, Phys. Rev. Lett.\n90, 146801 (2003).\n70B. A. Bernevig, J. Orenstein, and S. C. Zhang, Phys. Rev.\nLett. 97, 236601 (2006).\n71R. Winkler, Spin-Orbit Coupling E\u000bects in Two-\nDimensional Electron and Hole Systems (Springer, Berlin,\n2003).\n72J. D. Koralek, C. P. Weber, J. Orenstein, B. A. Bernevig,\nS.-C. Zhang, S. Mack, and D. D. Awschalom, Nature 458,\n610 (2009).\n73M. Duckheim, D. L. Maslov, and D. Loss, Phys. Rev. B\n80, 235327 (2009).\n74T. Meng, J. Klinovaja, and D. Loss, Phys. Rev. B 89,\n205133 (2014).\n75C. L. Kane, R. Mukhopadhyay, and T. C. Lubensky, Phys.\nRev. Lett. 88, 036401 (2002).\n76D. Poilblanc, G. Montambaux, M. H\u0013 eritier, and P. Lederer,\nPhys. Rev. Lett. 58, 270 (1987).\n77L. P. Gor'kov and A. G. Lebed, Phys. Rev. B 51, 3285\n(1995).\n78J. Klinovaja and D. Loss, Phys. Rev. Lett. 111, 196401\n(2013).\n79J. C. Y. Teo and C. L. Kane, Phys. Rev. B 89, 085101\n(2014).\n80J. Klinovaja and D. Loss, Eur. Phys. J. B 87, 171 (2014)\n81T. Meng, P. Stano, J. Klinovaja, and D. Loss, Eur. Phys.\nJ. B 87, 203 (2014).\n82T. Neupert, C. Chamon, C. Mudry, and R. Thomale, Phys.\nRev. B 90, 205101 (2014).\n83E. Sagi and Y. Oreg, Phys. Rev. B 90, 201102(R) (2014).\n84J. Klinovaja, Y. Tserkovnyak, and D. Loss, Phys. Rev. B\n91, 085426 (2015).\n85R. A. Santos, C.-W. Huang, Y. Gefen, and D. B. Gutman,\nPhys. Rev. B 91, 205141 (2015).\n86P.-H. Huang, J.-H. Chen, P. R. S. Gomes, T. Neupert, C.\nChamon, and C. Mudry, Phys. Rev. B 93, 205123 (2016).\n87P. Le Jeune, D. Robart, X. Marie, T. Amand, M. Brosseau,\nJ. Barrau, and V. Kalevcih, Semicond. Sci. Technol. 12,\n380 (1997).\n88A. Malinowski and R. T. Harley, Phys. Rev. B 62, 2051\n(2000).\n89M. A. Toloza Sandoval, A. Ferreira da Silva, E. A. de An-\ndrada e Silva, and G. C. La Rocca, Phys. Rev. B 86, 195302\n(2012).\n90See the Supplemental Material for (i) details of deriva-\ntion the topological criterion; (ii) an analytical description\nof helical edge modes and Majorana corner states; (iii) a\nstudy of the role of anisotropy on the topological phase\ndiagram; (iv) an explicit form of the discretized Hamil-\ntonian used in numerical calculations; (v) micromagneticsimulations of the magnetization dynamics in the driven\ntriple-layer setup.\n91The initial presence of the e\u000bective symmerty Te\u000bis not\ncrucial for our setup, as this symmetry is broken by the\nin-plane \feld anyway, thus, our setup is stable against dis-\norder and does not rely on the presence of particular spatial\nsymmetries.\n92R. Jackiw and C. Rebbi, Phys. Rev. D 13, 3398 (1976).\n93R. Jackiw and J. Schrie\u000ber, Nucl. Phys. B 190, 253 (1981).\n94W. Chang, S. M. Albrecht, T. S. Jespersen, F. Kuemmeth,\nP. Krogstrup, J. Nygard, and C. M. Marcus, Nat. Nan. 10,\n232 (2015).\n95Z. Wan, A. Kazakov, M. J. Manfra, L. N. Pfei\u000ber, K. W.\nWest, and L. P. Rokhinson, Nat. Comm. 6, 7426 (2015).\n96C. Reeg, D. Loss, and J. Klinovaja, Beilstein J. of Nan-\notechnol. 9, 1263 (2018).\n97C. Kittel, Phys. Rev. 73, 155 (1948).\n98C. Kittel, J. Phys. Radium 12, 3 (1951).\n99C. D. Stanciu and F. Hansteen, A. V. Kimel, A. Kirilyuk,\nA. Tsukamoto, A. Itoh, and T. Rasing, Phys. Rev. Lett.\n99, 047601 (2007).\n100A. Kirilyuk, A. V. Kimel, and T. Rasing, Rev. Mod. Phys.\n82, 2731 (2010).\n101S. Higashikawa, H. Fujita, M. Sato, arXiv:1810.01103\n(2018).\n102R.-C. Peng, J.-M. Hu, K. Momeni, J.-J. Wang, L.-Q.\nChen, and C.-W. Nan, Scienti\fc Reports 6, 27561 (2016).\n103Q. Wang, J. Domann, G. Yu, A. Barra, K. L. Wang, and\nG. P. Carman, Phys. Rev. Applied 10, 034052 (2018).\n104I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten,\nM. V. Costache, S. Au\u000bret, S. Bandiera, B. Rodmacq, A.\nSchuhl, and P. Gambardella. Nature 476, 189 (2011).\n105L. Liu, C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and\nR. A. Buhrman, Science 336, 6081 (2012).\n106P. Gambardella and I. M. Miron, Phil. Trans. R. Soc. A\n369, 3175 (2011).\n107J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back,\nand T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015).\n108A. Manchon, I. M. Miron, T. Jungwirth, J. Sinova, J.\nZelezn\u0013 y, A. Thiaville, K. Garello, and P. Gambardella,\narXiv:1801.09636 (2018).1\nSupplemental Material: Floquet Second-Order Topological Superconductor\nDriven via Ferromagnetic Resonance\nKirill Plekhanov,1Manisha Thakurathi,1Daniel Loss,1and Jelena Klinovaja1\n1Department of Physics, University of Basel, Klingelbergstrasse 82, CH-4056 Basel, Switzerland\nS1. Topological criterion in the isotropic limit\nIn the isotropic limit, \u000bx=\u000by, when the in-plane component of the oscillating magnetic \feld is zero, tk\nZ= 0, the\nsystem described by the Hamiltonian given in Eq. (4) has a continuous rotation symmetry in the xyplane. The\ncorresponding symmetry operator reads Urot\nJz(\u0012) = exp (\u0000i\u0012\u0011zJz=~), with\u0012being the rotation angle and Jzthez\ncomponent of the orbital angular momentum operator. Thus, the energy structure depends only on the absolute\nvalue of the momentum jkj=q\nk2x+k2yand can be calculated along a particular direction in the xyplane. For\nsimplicity we perform the calculations along the axis ky= 0 [see Fig. S1]. This results in the following Hamiltonian\ndensity\nHk(ky= 0) =\u0012~2k2\nx\n2m\u0000\u0016\u0013\n\u0011z\u0000\u000bxkx\u0011z\u001by+ \u0001 sc\u0011y\u001by+t?\nZ\u0011z\u001cx\u001bz+~!\u0011z\u0012\u001cz\u0000\u001c0\n2\u0013\n: (S1)\nHere, similar to the main text, we work in the Nambu basis \ty\nk= ( y\n1\"k, y\n1#k, y\n\u00161\"k, y\n\u00161#k, 1\"\u0000k, 1#\u0000k, \u00161\"\u0000k,\n \u00161#\u0000k). We de\fne by \u0011ithe Pauli matrices (the 2 \u00022 identity matrix for i= 0) acting on the particle-hole space, \u001ci{ on\nthe Floquet space, \u001bi{ on the space associated with the two spin components in the zdirection. The total Hamiltonian\nacts in the corresponding tensor-product space and for notational simplicity we suppress the explicit writing of the\ntensor product sign and the identity matrices. The system is characterized by an e\u000bective time-reversal symmetry\nTe\u000b=UTK=\u0000i\u001cz\u001byKwithUy\nTHkUT=H\u0003\n\u0000k, a particle-hole symmetry P=UPK=\u0011xKwithUy\nPHkUP=\u0000H\u0003\n\u0000k,\nand a chiral symmetry UC=PTe\u000b=\u0000i\u0011x\u001cz\u001bywithUy\nCHkUC=\u0000Hk. From this we deduce that the system belongs\nto the DIII symmetry class, characterized by Z2topological invariants, and has two distinct topological phases.\nIn the following, it will be convenient to change to the spin basis with quantization axis along the ydirection, using\nthe unitary rotation in the yzplane described by the operator Urot\n\u001bx= exp (\u0000i\u0019\u0011z\u001bx=4). This transformation satis\fes\n~\tk=Urot\n\u001bx\tkand ~Hk=Urot\n\u001bxHkUrot\n\u001bxy, with\n~Hk(ky= 0) =\u0012~2k2\nx\n2m\u0000\u0016\u0013\n\u0011z\u0000\u000bxkx\u001bz+ \u0001 sc\u0011y\u001by+t?\nZ\u001cx\u001by+~!\u0011z\u0012\u001cz\u0000\u001c0\n2\u0013\n: (S2)\nIn the new spin basis the Rashba SOI term acts on the spin component along the zaxis and we denote by \u001b= 1;\u00161\nits two possible orientations. In order to achieve the resonance condition, we \frst \fx the chemical potential \u0016to be\nsmaller in absolute value than the SOI energy, \u0000Eso<\u0016< 0. For instance, we restrict the discussion to the Floquet\nband\u001c= 1 only. As a result of the Rashba SOI, which lifts the spin degeneracy, there are four Fermi momenta\nkF\n1\u001b\u0006=\u001bkso\u0006k\u0016;withk\u0016=r\n2m(Eso+\u0016)\n~2; kso=m\u000bx\n~2;andEso=~2k2\nso\n2m: (S3)\nThe resonance condition is \fxed at the momentum kres=kF\n11\u0000, where the spin-split band \u001b= 1 crosses the chemical\npotential\u0016, as shown in Fig. S1. The resonance frequency is tuned to the energy di\u000berence between the two spin-split\nbands. Since the energy of the band \u001b= 1 atkresis zero, ~!is simply equal to the energy of the band \u001b=\u00161, which\ncan be written as\n~!=~2(kres+kso)2\n2m\u0000~2(kres\u0000kso)2\n2m= 4Eso \n1\u0000r\nEso+\u0016\nEso!\n: (S4)\nAccording to the Floquet formalism [1{4], the second Floquet band \u001c=\u00161 is shifted in energy with respect to the \frst\nband\u001c= 1 by an energy \u0000~!. Hence, it crosses the chemical potential \u0016at four momenta\nkF\n\u00161\u001b\u0006=\u001bkso\u0006(2kso\u0000k\u0016): (S5)\nWhen!is at resonance, the following identities hold true: kF\n\u00161\u00161+=kF\n11\u0000=kresandkF\n\u001611\u0000=kF\n1\u00161+=\u0000kres.2\nFIG. S1. The band structure of the 2DEG with SOI in the isotropic regime \u000bx=\u000byexpressed as a function of kxatky= 0\nand chemical potential \u0016(dashed line at bottom). The frequency !of a periodic drive is chosen resonantly such that the two\nFloquet bands (corresponding to \u001c= 1;\u00161) have the same smallest Fermi momentum kres(indicated by green circles) for the\ntwo spin-split bands (represented by the blue and red colors). Operators R\u001c\u001bandL\u001c\u001bcorrespond to the slowly varying left\nand right movers. The symbols colored in green indicate the operators involved in the resonant process at kres.\nTo see the e\u000bect of t?\nZand \u0001 scanalytically, we linearize the spectrum around the Fermi momenta kF\n\u001c\u001b\u0006and represent\nthe original operators in terms of slowly varying left and right moving \felds [5, 6] using the following relations\n~ 1\u001b(x) =~R1\u001b(x)e\u0000ikF\n1\u001b+x+~L1\u001b(x)e\u0000ikF\n1\u001b\u0000x;\n~ \u00161\u001b(x) =~R\u00161\u001b(x)e\u0000ikF\n\u00161\u001b+x+~L\u00161\u001b(x)e\u0000ikF\n\u00161\u001b\u0000x: (S6)\nIn the basis associated with the slowly varying \felds, the Hamiltonian density becomes\n~Hk(ky= 0) = ~vF\u000ek\u001az+ \u0001 sc\u001ax\u0011y\u001by+t?\nZ\n2\u001cx(\u001ax\u001by\u0000\u001ay\u001bx): (S7)\nHerevF=\u000bx=~is the Fermi velocity assumed to be equal for both Floquet bands, \u000ek=kx\u0000kF\n\u001c\u001b\u0006is the distance from\nthe Fermi momenta kF\n\u001c\u001b\u0006, and the Pauli matrices \u001aiact on the space of left and right movers. The bulk spectrum of\nthe linearized problem is given by\nE2\n1= (~vF\u000ek)2+ \u00012\nsc; E2\n2;\u0006= (~vF\u000ek)2+ (t?\nZ\u0006\u0001sc)2: (S8)\nBoth the Floquet-Zeeman and the superconducting terms induce an opening of the gap at the Fermi momenta \u0006kres,\nwhich leads to a competition between the two terms. In contrast, the gap at the remaining Fermi momenta is\nopened only by the superconducting term. When the Floquet amplitude t?\nZbecomes of the same strength as the\nsuperconducting pairing amplitude \u0001 sc, we observe the closing of the gap, indicating a topological phase transition.\nWhen \u0001 sc>t?\nZthe system is in the topologically trivial phase, while in the regime \u0001 sc/radicalbig\n/planckover2pi1γµ0(MS+Hs)ǫex. Under this condition, there are\nlower energy states than a uniformly magnetized state\neven atHext= 0, making resonant states still possible.\nThis can be understood by reminding the fact that spi-\nral states are a ground state rather than a ferromagnetic\nstate for systems accompanying a large DMI.\nWe now turn to the discussion about the linewidth\nin the out-of-plane geometry of applied magnetic fields.\nFigure 5 is the linewidth for different tilted angles of the\nexternal magnetic field. Because we consider a strong\nDMI case ( ǫDM≃ǫex), the peaks appear in the linewidth\ncurves. We find that the peak positions are nearly un-\nchanged with the tilting angle of magnetic field. In the\nrange of high frequency apart from the peak structures,\nthereisacriticalfrequencyabovewhichthelinewidth be-\ncomes zero, a similar behavior found in DMI-free cases25.\nIn Fig. 5(b), we plot the linewidth as a function of\nthe tilting angle when the FMR frequency is fixed. For\na small DMI case, the critical angle above which the\nlinewidth is zero is determined by Eq. (B2) and is simi-6\n0 10 20 30\nf (GHz)04080120∆H(2) (G)(a)\nθH (deg.)\n 0 \n60 \n70 \n80 \n0 10 20 30\nf (GHz)04080120∆H(2) (G)(a)\nθH (deg.)\n 0 \n60 \n70 \n80 \n0 20 40 60 80\nθH (deg.)0102030∆H(2) (G)(b)\nf (GHz)\n7\n15\n20\n30\n0 20 40 60 80\nθH (deg.)0102030∆H(2) (G)(b)\nf (GHz)\n7\n15\n20\n30\nFIG. 5. (color online) For a given DMI strength D= 1.5\nmJ/m2, we compare ∆ H(2)for different angles of applied\nmagnetic field. Other parameters are the same as those in\nFig. 2.\nlar to that in the work of Ref. [25]. On the other hand,\nthe critical angle in a strong DMI case is determined by\nEq. (B3).\nV. CONCLUSION\nIn this work, we have developed the theory of\ntwo-magnon scattering for a thin ferromagnetic film\nwith DMI as well as dipole-dipole, exchange, surface\nanisotropy, and Zeeman energies. In a quantum mechan-\nical way, we derive a expression of the FMR linewidth,\ntaking into account scattering from structural inhomo-\ngeneity, for example, DMI fluctuation induced by mi-\ncroscopic structural imperfection. We present the ex-\ntrinsic FMR linewidth in terms of various energy con-\ntributions, Eq. (21). In the presence of the DMI term,\nthe magnon dispersion exhibits rich resonant states, es-\npecially in small external magnetic fields. Furthermore,\ndue to the competition between the exchange and DMI\nenergies, the resonant states are formed at large kpoints\n(namely kd≃1). Different from the DMI-free case\n(kd≪1), it makes difficult to derive analytical expres-\nsions from the FMR linewidth.\nWe find that the characteristic linewidth broadening\nwith the DMI are twofold. One is the appearance of\npeak in the low frequency range when the DMI is strong.\nThis usually occurs when twolobes of the resonantstates\ntouch each other in magnon dispersion curves. The other\nis a finite value of the FMR linewidth at zero magneticfield. This may be not possible if a uniform ferromag-\nnetic system such as DMI-free case is the ground state.\nHowever, in the case of strong DMI, a non-colinear spin\nstate like spiral configuration has a lower energy than a\nferromagnetic one. Consequently, in the presence of the\nstrong DMI, there are still resonant states even at zero\nmagnetic field and thus a uniform spin state excited by\nmicrowaves in FMR experiments is still scattered into\nresonant states by inhomogeneities scattering.\nVI. ACKNOWLEDGEMENTS\nWe acknowledge R. D. McMichael for a fruitful discus-\nsion. This work was supported by the National Research\nFoundation of Korea (NRF-2015M3D1A1070465, NRF-\n2017R1A2B2006119) and KU-KIST Graduate School of\nConverging Science and Technology Program.\nAppendix A: DMI perturbation potential\nLet’s consider fluctuation from a Rashba-type spin-\norbit interaction51;\n∆ED=4/planckover2pi1γ\nMS/integraldisplay\ndrDd(r)/bracketleftbigg\nˆz·m×∂m\n∂x−ˆx·m×∂m\n∂z/bracketrightbigg\nwhereDd(r) is local DMI strength for each defect.\nOneoffeasiblemodelsforthe fluctuationofthisenergy\nisarandomspatialvariationofthe parameter Dd(r) with\nkeeping the magnon propagating vector m. This case\nmay occur structural defects whose characteristic size is\nlarger than or equal to the film thickness dwith weak\nDMI. Then, the fluctuation energy becomes, in Fourier\nspace,\n∆ED=/summationdisplay\nk,k′Vd(k−k′)ik′\nxd[m(−k)×m(k′)]·ˆz\n−/summationdisplay\nk,k′Vd(k−k′)ik′\nzd[m(−k)×m(k′)]·ˆx,(A1)\nwhere the defect potential energy is defined by,\nVd(r) =4/planckover2pi1γ\nMSdDd(r) =/summationdisplay\nkVd(k)eik·r.\nAs another model, we can take account of an abrupt\nchange of the magnetization vectors, for example, at\nedges of terrace on sample surfaces, near magnetic and\nnon-magnetic impurities, etc. Moreover, it is also known\nthat the tilt magnetization state (spiral or skyrmion\nphases) rather than a ferromagnetic state is more sta-\nble for a strong DMI strength.76For those cases, we ex-\npect that magnon modes are additionally modulated as\nˆm(r)ei(k+q)·rnear the defect ( qis a pitch vector, for ex-\nample, in a spiral state). Then, the fluctuation energy\nbecomes similar to Eq. (A1), but replacing kx→kx+qx\nandkz→kz+qz. By considering small defects with a7\ncharacteristic length λc(order of an impurity size ≪d),\nwe further set kx→1/λcandkz→1/λcin Eq. (A1).\nWe assume that the defect potential energy Vd(r) is\ncontributed from atom-like and Gaussian-shaped func-\ntions located at random sites;\nVd(r) =/summationdisplay\njva(r−Rj), va(r) =4/planckover2pi1γD\nMSde−r2/λ2\nimp(A2)\nwith a defect size, λimp. Then, by inserting Eq. (A2)\ninto Eq. (A1), the fluctuation energy becomes, in the\nlocal coordinates,\n∆ED=/summationdisplay\nk,k′Vd(k−k′)i(kx+k′\nx+2/λc)d\n2cosθM\nˆx3·[m(−k)×m(k′)].\n=/summationdisplay\nk,k′/summationdisplay\nκ,κ′=1,2mκ(−k)Uk,k′\nκκ′mκ′(k′) (A3)\nwhere\nUk,k′\n11=Uk,k′\n22= 0,\nUk,k′\n12=−Uk′,k\n21=/summationdisplay\nje−i(k−k′)·RjUa(k,k′).\nHere,Ua(k,k′) is given by,\nUa(k,k′) =2/planckover2pi1γD\nMSdfR(k−k′)W(k,k′)cosθM(A4)\nwhereW(k,k′) =i(k′\nx+kx)d/2 forλc/greaterorsimilard, whereas\nW=d/λcforλc≪d. Here,fR(k) =e−k2λ2\nimpπλ2\nimp/L2\nis a form factor of the atomic potential ( Lis a side length\nof the ferromagnetic film).\nAppendix B: Critical angle of external magnetic\nfields\nAbove a certain angle of external magnetic field, the\nlinewidth of FMR becomes zero. The critical angle is\ndetermined by demanding zero resonant states, or tiny\nsize of lobes, in the FMR condition of ǫ(k) =/planckover2pi1ωFMR. By\nrequiring a small value of k, it is straightforward to find\nkR(φk)d=/parenleftBig\nC1\n2C2/parenrightBig2\n+C0\nC2−/parenleftBig\nC1\n2C2+sinφk/parenrightBig2\nD2/C2(B1)\nwhere\nC0=ǫMs/bracketleftbig\nǫ0\n22cos2θM−(ǫ0\n11+ǫ0\n22)sin2θM/bracketrightbig\nC1= 2/radicalBig\nǫ0\n11ǫ0\n22ǫDMcosθM,\nC2=ǫMs/bracketleftbig\nǫ0\n11−ǫ0\n22sin2θM/bracketrightbig\n,\nD2= (ǫ0\n11+ǫ0\n22)ǫex+ǫ2\nMssin2θM(sin2θM−cos2θM)\n−1\n2sin2φkcos2θM[2ǫ2\nDM+ǫ2\nMs(1+cos2 θM)].Thus, for a small DMI ofC1\n2C2≤1, the critical angle is\ndetermined by solving the equation,\n/parenleftbiggC1\n2C2/parenrightbigg2\n+C0\nC2= 0 (B2)\nwhile,C1\n2C2>1,\n/parenleftbiggC1\n2C2/parenrightbigg2\n+C0\nC2−/parenleftbiggC1\n2C2−1/parenrightbigg2\n= 0.(B3)\nAppendix C: Evaluation of the linewidth function\nΓ(ω)\nThe linewidth function Γ( ωFMR) can be rewritten as,\nfrom Eq. (21),\nΓ(ω) =1\ncos(θM−θH)ǫex\n(ǫ0\n22+ǫ0\n11)πd2\nL2/summationdisplay\nkfc(k)\nǫ0\n11ǫ11(k)+ǫ0\n22ǫ22(k)/radicalbig\n4ǫ11(k)ǫ22(k)−[ǫ12(k)+ǫ21(k)]2δ(/planckover2pi1ω−ǫ(k)).(C1)\nFor a weak DMI ofC1\n2C2≤1 and due to kd≪1, the\ndelta function in the equation can be approximated by,\nδ(/planckover2pi1ω−ǫ(k))≃2/radicalbig\nǫ0\n11ǫ0\n22\n|C0|dδ(k−kR(φk))\nfrom Eq. (B1) and then, Γ( ω) is further simplified to\nΓ(ω)≃ǫex([ǫ0\n22]2+[ǫ0\n11]2)\n2πcos(θM−θH)(ǫ0\n22+ǫ0\n11)|C0|/integraldisplayφb\nφadφkR(φ)d\n(C2)\nwhere\nsinφa=−C1\n2C2−/radicalBigg/parenleftbiggC1\n2C2/parenrightbigg2\n+C0\nC2\nsinφb=−C1\n2C2+/radicalBigg/parenleftbiggC1\n2C2/parenrightbigg2\n+C0\nC2.\nNow we consider a strong DMI ofC1\n2C2≫1 where the\ndipole energy ǫMsis relatively unimportant. In this case,\neigenenergy of magnon can be approximated by, from\nEq. (7),\nǫ(k) = Im[ǫ12(k)]+/radicalBig\n(ǫ0\n11+ǫexk2d2)(ǫ0\n22+ǫexk2d2).\nBy solving this, one can show that the resonant states\nappear at,\nkR(φk)d≃ǫDM|sinφk|cosθM\nǫex,−π\n2≤φk≤0.(C3)\nThen, magnon density of states at a kpoint is given by,\nδ(/planckover2pi1ω−ǫ(k))≃1\nǫ0\n11+ǫ0\n22−/planckover2pi1ω+2˜ǫ2\nDM\nǫexδ(kd−kR(φk)d)8\nwhere we abbreviate ˜ ǫDM=ǫDM|sinφk|cosθM, and the associated linewidth function Γ( ω) becomes\nΓ(ω)≃1\n4πcos(θM−θH)\n/integraldisplay0\n−π/2dφkǫex\nǫ0\n11+ǫ0\n22−/planckover2pi1ω+2˜ǫ2\nDM\nǫex.(C4)\n∗kjlee@korea.ac.kr\n1V. Kambersk´ y, Czech. J. Phys. 26, 1366 (1976).\n2K. Gilmore, Y. U. Idzerda, and M. D. Stiles, Phys. Rev.\nLett.99, 027204 (2007).\n3V. Kambersk´ y, Phys. Rev. B 76, 134416 (1970).\n4I. Garate, K. Gilmore, M. D. Stiles, and A. H. MacDonald,\nPhys. Rev. B 79, 104416 (2009).\n5H. Ebert, S. Mankovsky, D. Kodderitzsch, and P. J. Kelly,\nPhys. Rev. Lett. 107, 066603 (2011).\n6Z. Yuan, K. M.D. Hals, Y. Liu, A. A. Starikov, A. Brataas,\nand . J. Kelly, Phys. Rev. Lett. 113, 266603 (2014).\n7M. A. W. Schoen, D. Thonig, M. L. Schneider, T. J. Silva,\nH. T. Nembach, O. Eriksson, O. Karis, and J. M. Shaw,\nNat. Phys. 12, 839 (2016).\n8J. Z. Sun, Phys. Rev. B 62, 570 (2000).\n9S.-W. Lee and K.-J. Lee, Proc. IEEE 104, 1831 (2016).\n10N. Smithand P. Arnett, Appl. Phys. Lett. 78, 1448 (2001).\n11V. L. Safonov and H. N. Bertram, Phys. Rev. B 65, 172417\n(2002).\n12S. Mizukami, Y. Ando, and T. Miyazaki, J. Magn. Magn.\nMater.226, 1640 (2001).\n13K. Kobayashi, N. Inaba, N. Fujita, Y. Sudo, T. Tanaka, M.\nOhtake, M. Futamoto, and F. Kirino, IEEE Trans. Magn.\n45, 2541 (2009).\n14M. C. Langner, C. L. S. Kantner, Y. H. Chu, L. M. Martin,\nP. Yu, J. Seidel, R. Ramesh, and J. Orenstein, Phys. Rev.\nLett.102, 177601 (2009).\n15X. Liu, W. Zhang, M. J. Carter, and G. Xiao, J. Appl.\nPhys.110, 033910 (2011).\n16T. Weindler, H. G. Bauer, R. Islinger, B. Boehm, J. -Y.\nChauleau, and C. H. Back, Phys. Rev. Lett. 113, 237204\n(2014).\n17M. Sparks, R. Loudon, and C. Kittel, Phys. Rev. 122, 791\n(1961).\n18Z.CelinskiandB.Heinrich, J.Appl.Phys. 70, 5935(1991).\n19M. J. Hurben, D. R. Franklin, and C. E. Patton, J. Appl.\nPhys.81, 7458 (1997).\n20R. D. McMichael, M. D. Stiles, P. J. Chen, and W. F.\nEgelhoff, J. Appl. Phys. 83, 7037 (1998).\n21R. Arias and D. L. Mills, Phys. Rev. B. 60, 7395 (1999).\n22R. D. McMichael, D. J. Twisselmann, and A. Kunz, Phys.\nRev. Lett. 90, 227601 (2003).\n23R. D. McMichael and P. Krivosik, IEEE Trans. Magn. 40,\n2 (2004).\n24Kh. Zakeri, J. Lindner, I. Barsukov, R. Meckenstock, M.\nFarle, U. von H¨ orsten, H. Wende, W. Keune, J. Rocker, S.\nS. Kalarickal, K. Lenz, W. Kuch, K. Baberschke, and Z.\nFrait, Phys. Rev. B. 76, 104416 (2007).\n25P. Landeros, R. E. Arias, and D. L. Mills, Phys. Rev. B.\n77, 214405 (2008).\n26I.M. Miron, K.Garello, G.Gaudin, P.-J. Zermatten, M.V.\nCostache, S. Auffret, S. Bandiera, B. Rodmacq, A. Schuhl,and P. Gambadella, Nature 476, 189 (2011).\n27L. Liu, C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and\nR. A. Buhrman, Science 336, 555 (2012).\n28S. Emori, U. Bauer, S.-M. Ahn, E. Martinez, and G. S. D.\nBeach, Nat. Mater. 12, 611 (2013).\n29K.-S. Ryu, L. Thomas, S.-H. Yang, and S. Parkin, Nat.\nNanotechnol. 8, 527 (2013).\n30P. M. Haney, H.-W. Lee, K.-J. Lee, A. Manchon, and M.\nD. Stiles, Phys. Rev. B 87, 174411 (2013).\n31P. M. Haney, H.-W. Lee, K.-J. Lee, A. Manchon, and M.\nD. Stiles, Phys. Rev. B 88, 214417 (2013).\n32X. Fan, H. Celik, J. Wu, C. Ni, K.-J. Lee, V. O. Lorenz,\nand J. Q. Xiao, Nat. Commun. 5, 3042 (2014).\n33X. Qiu, K. Narayanapillai, Y. Wu, P. Deorani, D.-H. Yang,\nW.-S. Noh, J.-H. Park, K.-J. Lee, H.-W. Lee, and H. Yang,\nNat. Nanotechnol. 10, 333 (2015).\n34Y.-W. Oh, S. Chris Baek, Y. M. Kim, H. Y. Lee, K.-D.\nLee, C.-G. Yang, E.-S. Park, K.-S. Lee, K.-W. Kim, G.\nGo, J.-R. Jeong, B.-C. Min, H.-W. Lee, K.-J. Lee, and\nB.-G. Park, Nat. Nanotechnol. 11, 878 (2016).\n35V. P. Amin and M. D. Stiles, Phys. Rev. B 94, 104419\n(2016).\n36V. P. Amin and M. D. Stiles, Phys. Rev. B 94, 104420\n(2016).\n37A. M. Humphries, T. Wang, E. R. J. Edwards, S. R. Allen,\nJ. M. Shaw, H. T. Nembach, J. Q. Xiao, T. J. Silva, and\nX. Fan, Nat. Commun. 8, 911 (2017).\n38S. C. Baek, V. P. Amin, Y.-W. Oh, G. Go, S.-J. Lee, M. D.\nStiles, B.-G. Park, and K.-J. Lee, arXiv:1708.06864 (2017) .\n39K.-W. Kim, K.-J. Lee, J. Sinova, H.-W. Lee, and M. D.\nStiles, Phys. Rev. B 96, 104438 (2017).\n40I. E. Dzyaloshinskii, Sov. Phys. JETP 5, 1259 (1957).\n41T. Moriya, Phys. Rev. 120, 91 (1960).\n42U. K. R¨ oßler, A. N. Bogdanov, and C. Pfleiderer, Nature\n(London) 442, 797 (2006).\n43M. Uchida, Y. Onose, Y. Matsui, and Y. Tokura, Science\n311, 359 (2006).\n44S. X.HuangandC. L. Chien, Phys.Rev. Lett. 108, 267201\n(2012).\n45A. Fert and P. M. Levy, Phys. Rev. Lett. 44, 1538 (1980).\n46M. Bode, M. Heide, K. von Bergmann, P. Ferriani, S.\nHeinze, G. Bihlmayer, A. Kubetzka, O. Pietzsch, S. Bl¨ ugel,\nand R. Wiesendanger, Nature (London) 447, 190 (2007).\n47M. Heide, G. Bihlmayer, and S. Bl¨ ugel, Phys. Rev. B 78,\n140403(R) (2008).\n48L. Udvardi and L. Szunyogh, Phys. Rev. Lett. 102, 207204\n(2009).\n49Kh. Zakeri, Y. Zhang, J. Prokop, T. H. Chuang, N. Sakr,\nW. X. Tang, and J. Kirschner, Phys. Rev. Lett. 104,\n137203 (2010).\n50A. T. Costa, R. B. Muniz, S. Lounis, A. B. Klautau, and\nD. L. Mills, Phys. Rev. B 82, 014428 (2010).9\n51K.-W. Kim, H.-W. Lee, K.-J. Lee, and M. D. Stiles, Phys.\nRev. Lett. 111, 216601 (2013).\n52A. Thiaville, S. Rohart, ´E. Ju´ e, V. Cros, and A. Fert, Eu-\nrophys. Lett. 100, 57002 (2012).\n53G. Chen, J. Zhu, A. Quesada, J. Li, A. T. N’Diaye, Y.\nHuo, T. P. Ma, Y. Chen, H. Y. Kwon, C. Won, Z. Q. Qiu,\nA. K. Schmid, and Y. Z. Wu, Phys. Rev. Lett. 110, 177204\n(2013).\n54S. M¨ ulbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch,\nA. Neubauer, R. Georgii, and P. B¨ oni, Science 323, 915\n(2009).\n55X. Z. Yu, Y. Onose, N. Kanazawa, J. H. park, J. H. Han,\nY. Matsui, N. Nagaosa, and Y. Tokura, Nature (London)\n465, 901 (2010).\n56W. Jiang, P. Upadhyaya, W. Zhang, G. Yu, M. B.\nJungfleisch, F. Y. Fradin, J. E. Pearson, Y. Tserkovnyak,\nK. L. Wang, O. Heinonen, S. G. E. te Velthuis, and A.\nHoffmann, Science 349, 283 (2015).\n57S. Woo, K. Litzius, B. Kr¨ uger, M.-Y. Im, L. Caretta, K.\nRichter, M. Mann, A. Krone, R. M. Reeve, M. Weigand, P.\nAgrawal, I. Lemesh, M.-A. Mawass, P. Fischer, M. Kl¨ aui,\nand G. S. D. Beach, Nat. Mater. 15, 501 (2016).\n58C. Moreau-Luchaire, C. Moutas, N. Reyren, J. Sampaio,\nC. A. F. Vaz, N. Van Horne, K. Bouzehouane, K. Garcia,\nC. Deranlot, P. Warnicke, P. Wohlhter, J.-M. George, M.\nWeigand, J. Raabe, V. Cros, and A. Fert, Nat. Nanotech-\nnol.11, 444 (2016).\n59O.Boulle, J. Vogel, H.Yang, S.Pizzini, D.deS.Chaves, A.\nLocatelli, T. O. Mentes, A. Sala, L. D. Buda-Prejbeanu, O.\nKlein, M. Belmeguenai, Y. Roussigne, A. Stashkevich, S.\nM. Cherif, L. Aballe, M. Foerster, M. Chshiev, S. Auffret,\nI. M. Miron, and G. Gaudin, Nat. Nanotechnol. 11, 449\n(2016).\n60S. D. Pollard, J. A. Garlow, J. Yu, Z. Wang, Y. Zhu, and\nH. Yang, Nat. Commun. 8, 14761 (2017).61D. Cort´ es-Ortu˜ no and P. Landeros, J. Phys.: Condens.\nMatter25, 156001 (2013).\n62J.-H. Moon, S.-M. Seo, K.-J. Lee, K.-W. Kim, J. Ryu, H.-\nW. Lee, R. D. McMichael, and M. D. Stiles, Phys. Rev. B\n88, 184404 (2013).\n63K. Di, V. L. Zhang, H. S. Lim, S. C. Ng, M. H. Kuok, J.\nYu, J. Yoon, X. Qiu, and H. Yang, Phys. Rev. Lett. 114,\n047201 (2015).\n64H. T. Nembach, J. M. Shaw, M. Weiler, E. Ju´ e, and T. J.\nSilva, Nat. Phys. 11, 825 (2015).\n65J. Cho, N.-H. Kim, S. Lee, J.-S. Kim, R. Lavrijsen, A.\nSolignac, Y. Yin, D.-S. Han, N. J. J. van Hoof, H. J.\nM. Swagten, and B. Koopmans, Nat. Commun. 6, 7635\n(2015).\n66J. M. Lee, C. Jang, B.-C. Min, S.-W. Lee, K.-J. Lee, and\nJ. Chang, Nano Lett. 16, 62 (2016).\n67S. Seki, Y. Okamura, K. Kondou, K. Shibata, M. Kubota,\nR. Takagi, F. Kagawa, M. Kawasaki, G. Tatara, Y. Otani,\nand Y. Tokura, Phys. Rev. B 93, 235131 (2016).\n68J.-V. Kim, R. L. Stamps, and R. E. Camley, Phys. Rev.\nLett.117, 197204 (2016).\n69J. Lan, W. Yu, R. Wu, andJ. Xiao, Phys.Rev. X 5, 041049\n(2015).\n70W. Yu, J. Lan, R. Wu, and J. Xiao, Phys. Rev. B 94,\n140410(R) (2016).\n71J. Lan, W. Yu, and J. Xiao, Nat. Commun. 8, 178 (2017).\n72S.-J.Lee, J.-H.Moon, H.-W.Lee, andK.-J.Lee, submitted\n(2017).\n73R. W. Damon and J. R. Eshbach, J. Phys. Chem. Solids\n19, 308 (1961).\n74T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940).\n75B. A. Kalinikos, J. Phys. C, Solid State Phys. 19, 7013\n(1986).\n76J. Iwasaki, M. Mochizuki, and N. Nagaosa, Nat. Commun.\n4, 1463 (2009)." }, { "title": "2301.03256v1.X_ray_detected_ferromagnetic_resonance_techniques_for_the_study_of_magnetization_dynamics.pdf", "content": "X-ray detected ferromagnetic resonance techniques\nfor the study of magnetization dynamics\nGerrit van der Laan1and Thorsten Hesjedal2\n1Diamond Light Source, Harwell Science and Innovation Campus, Didcot, OX11 0DE, United Kingdom\n2Department of Physics, Clarendon Laboratory, University of Oxford, Oxford, OX1 3PU, United Kingdom\n(Dated: January 10, 2023)\nElement-speci\fc spectroscopies using synchrotron-radiation can provide unique insights into\nmaterials properties. The recently developed technique of X-ray detected ferromagnetic resonance\n(XFMR) allows studying the magnetization dynamics of magnetic spin structures. Magnetic\nsensitivity in XFMR is obtained from the X-ray magnetic circular dichroism (XMCD) e\u000bect, where\nthe phase of the magnetization precession of each magnetic layer with respect to the exciting\nradio frequency is obtained using stroboscopic probing of the spin precession. Measurement of\nboth amplitude and phase response in the magnetic layers as a function of bias \feld can give a\nclear signature of spin-transfer torque (STT) coupling between ferromagnetic layers due to spin\npumping. Over the last few years, there have been new developments utilizing X-ray scattering\ntechniques to reveal the precessional magnetization dynamics of ordered spin structures in the GHz\nfrequency range. The techniques of di\u000braction and re\rectometry ferromagnetic resonance (DFMR\nand RFMR) provide novel ways for the probing of the dynamics of chiral and multilayered magnetic\nmaterials, thereby opening up new pathways for the development of high-density and low-energy\nconsumption data processing solutions.\nKeywords: FMR, XMCD, X-ray scattering, X-ray re\rectivity, spin structures\nI. INTRODUCTION\nMagnetization dynamics is at the heart of high fre-\nquency magnetic nanoscale devices based on spin waves,\nspin pumping, and spin-torque oscillators in the GHz\nfrequency range. Traditionally, ferromagnetic resonance\n(FMR) has been a work horse technique to determine the\nfundamental parameters for magnetic resonance and re-\nlaxation in thin \flms. The recent growing complexity of\nmany modern magnetic materials and devices requires\nthe development of advanced measurement techniques\nthat more directly reveal the microscopic origin of the\ndynamical magnetic interactions that are at play.\nThe novel techniques of X-ray detected FMR (XFMR)\nenables studying the magnetization dynamics of indi-\nvidual layers, where element-speci\fc magnetic contrast\nis obtained using the X-ray magnetic circular dichroism\n(XMCD) e\u000bect [1]. Not only can the FMR signal be mon-\nitored in X-ray absorption, it can also be done in X-ray\ndi\u000braction and re\rectivity, using techniques termed as\nDFMR and RFMR, respectively [2]. In these X-ray mea-\nsurements, time-resolved FMR gives both the amplitude\nand phase of the spin precession for the di\u000berent chemical\nelements, and hence di\u000berent layers, in the sample. The\nchallenge of such measurements lays in the fact that the\nprecession cone angle is small ( <1\u000e) and that the preces-\nsion frequency is on the order of GHz. The solution is\nto use lock-in techniques and to detect the phase of the\nprecession stroboscopically by using the time structure\nof the X-ray pulses from the synchrotron ( \u0018500 MHz,\ni.e., corresponding to a period between the pulses of 2\nns). The radio frequency (RF) \feld applied to drive the\nspin precession is synchronized with the X-ray pulses us-\ning the clock of the synchrotron. Therefore, each X-raypulse measures the magnetization cone at precisely the\nsame point in the precession cycle. Hence, XFMR com-\nbines the techniques of FMR and XMCD. Thus, the spin\nprecession along the bias \feld is pumped by the RF \feld\nto generate the magnetic resonance (i.e., FMR), whose\namplitude and phase is probed by the synchronized X-\nray pulses using the XMCD e\u000bect. The time dependence\nis recorded using a delay line to vary the phase of the RF\nsignal with respect to the X-ray pulses.\nDuring the last few years, many XFMR studies either\nin time-averaged or time-resolved mode have been re-\nported [3{52]. The \frst element-speci\fc and time-depen-\ndent measurement of the magnetization dynamics using\npump-probe XMCD was reported by Bailey et al. [3] on a\npermalloy (Py = Ni 80Fe20) thin \flm, where the moments\non the Ni and Fe sites were found to precess together at\nall frequencies, and by Arena et al. [4] on a Py/Cu/CoZr\ntrilayer, where at resonance, a weak ferromagnetic cou-\npling was found in the phase and amplitude response of\nindividual layers across resonance.\nThe amplitude and phase response of the magnetic\nprobe layer measured by XFMR provides a signature for\neither static exchange interaction in strongly exchange-\ncoupled bilayers or spin-transfer torque (STT) coupling\ndue to spin pumping. Marcham et al. [25] \frst evidenced\nSTT in a CoFe/Cu/Py spin valve using XFMR where the\n\feld dependence of the \fxed layer phase showed a clear\nsignature of STT due to spin pumping. Using XFMR,\nBaker et al. [28] reported a strong anisotropy of the spin\npumping, providing new opportunities for device appli-\ncations.\nPreviously, time-resolved XFMR has been reviewed in\ngreat detail in Ref. [1]. Here, we present a timely up-\ndate, especially emphasizing the newly developed time-arXiv:2301.03256v1 [cond-mat.mtrl-sci] 9 Jan 20232\nresolved FMR techniques in X-ray re\rectivity and di\u000brac-\ntion.\nThe outline of the remainder of this paper is as follows.\nSec. II gives a brief theoretical background of magnetiza-\ntion dynamics and STT. Sec. III describes the experimen-\ntal setup, conditions, and considerations for the various\nXFMR techniques. Sec. IV highlights several recent ex-\namples of XFMR, DFMR, and RFMR experiments and\nmentions their scienti\fc impact. Finally, conclusions are\ndrawn in Sec. V.\nII. BACKGROUND ON FMR AND STT\nA. Ferromagnetic resonance (FMR)\nBefore presenting the experimental details and show-\ncasing several recent examples, we will brie\ry introduce\nsome relevant background material.\nFMR arises when the energy levels of a quantized sys-\ntem of electronic moments are Zeeman split by a uniform\nmagnetic \feld and the system absorbs energy from an os-\ncillating magnetic \feld [53]. A resonance occurs when the\ntransverse AC \feld is applied at the Larmor frequency\ncorresponding to the energy di\u000berence between the mag-\nnetic levels, i.e., ~!= \u0001E. The spin precession in a\nsingle-domain magnetic material can be described with\nthe equation of motion, the so-called Landau-Lifshitz-\nGilbert (LLG) equation,\n_ m=\u0000\rm\u0002He\u000b+\u000b(m\u0002_ m); (1)\nwhere the e\u000bective \feld He\u000b=\u0000@F(M)=@Mis ob-\ntained by minimization of the free energy Fwith re-\nspect to the magnetization M. The free energy contains\nterms such as the exchange, Dzyaloshinskii-Moriya, de-\nmagnetization, magnetocrystalline anisotropy, magneto-\nstatic, external Zeeman \feld, and elastic energy. Further,\n_ m=\u000em=\u000et; the reduced magnetization is m=M=Ms,\nwhereMs=jmjis the saturation magnetization; and\n\r=g\u0016B=~is the gyromagnetic ratio, where \u0016Bis the\nBohr magneton and gis the Land\u0013 e (spectroscopic split-\nting) g-factor. The dimensionless damping parameter\n\u000b\u001c1 (typically 10\u00003{10\u00002for 3dmetals) determines\nthe width of the resonance absorption peak.\nThe \frst right-hand term in Eq. (1) corresponds to\nthe torque due to the e\u000bective \feld He\u000b. In a classical\npicture,\u001c=dS=dtequates to the time change in an-\ngular momentum S, which leads to the spin precession.\nThe second right-hand term corresponds to the damp-\ning, which can also be written in the form of the Gilbert\ndamping term\u0000\u000b\r(m\u0002m\u0002He\u000b). Both torque and\ndamping are vectorially sketched in Fig. 1(a). Without\nexternal RF excitation, the magnetization would relax to\nthe steady state given by Brown's equation, m\u0002He\u000b= 0.\nLinearization of the LLG equation gives the relation\nbetween the frequency \u00170(or circular frequency !0) and\n\feld, which in the form of the Kittel equation is writtenas [1]\n2\u0019\u00170\u0011!0=\rp\nHe\u000bBe\u000b=\rp\nHe\u000b(Ms+He\u000b):(2)\nB. Spin-transfer torque (STT)\nThe layer selectivity of XFMR makes this technique\na unique probe to investigate STT and related spin cur-\nrents in multi-layered spin valves [1]. STT is the e\u000bect\nin which the spin direction in a magnetic layer can be\nmodi\fed using a spin-polarized current [54, 55].\nSpin pumping occurs when the precessing magnetiza-\ntion vector generated by FMR in a ferromagnetic (FM)\nlayer emits a pure spin current into an adjacent normal\nmetal (NM) layer [56]. Traditionally, spin currents have\nbeen probed using indirect measurements. For instance,\nin the metals through which they \row they can create an\nelectrical voltage drop perpendicular to the spin current\ndirection, or a torque that bends the magnetization di-\nrection. However, such indirect measurements are often\nambiguous because they are also in\ruenced by other fac-\ntors, such as magnetic proximity e\u000bects at the interface.\nSTT gives an extra term in the LLG equation, which\nis (anti)-parallel to the (anti)-damping (see Fig. 1(a)).\nAccording to Slonczewski [54], the adiabatic torque is\n\u001cs=\u000bsm\u0002_ m, where\u000bsis the STT damping. The\nspin current pumped across a FM/NM interface due to\nprecession is [56]\nIs=~\n4\u0019g\"#\ne\u000bm\u0002_ m; (3)\nwhereg\"#\ne\u000bis the e\u000bective spin-mixing conductance. The\nspin pumping depends critically on the FM/NM inter-\nface (the material-dependent g\"#\ne\u000b) and the spin di\u000busion\nlength in the NM.\nFor two FM layers iandjwith di\u000berent resonance fre-\nquencies and coupled by both spin pumping (dynamic ex-\nchange coupling) and static exchange coupling, the cou-\npled LLG equations are\n_ mi=\u0000\rmi\u0002He\u000b;i+\u000b0\nimi\u0002_ mi\n+\u000bs\nimi\u0002(_ mi\u0000_ mj) +Aexmj\u0001mi;(4)\nand equivalently when exchanging i$j, where miis\nthe magnetization direction, He\u000b;ithe e\u000bective \feld, \u000b0\ni\nthe Gilbert damping, and \u000bs\nithe STT damping in layer\ni. The spin pumping induced coupling is determined by\n\u000bs\niand the static exchange coupling by Aex.\nIII. EXPERIMENTAL\nXFMR provides an element-speci\fc and time-resolved\nmeasurement of the precessional dynamics of each FM\nlayer on a ps time scale, where the spin precession in-\nduced by a driving RF signal is detected using the XMCD3\nHeffh(t) m(t)Sample\nCPW(b)\nm×Heffz\nyHeff\nm(t)m ×m ×Heff\nSTT\nx(a)\nFIG. 1. (a) Precession, damping, and spin transfer torque\n(STT) in FMR. The precession m\u0002He\u000baround the e\u000bective\n\feldHe\u000bis damped by the Gilbert term m\u0002m\u0002He\u000b. The\nspin-transfer torque is parallel (antiparallel) to the Gilbert\ndamping term, and can enhance (oppose) the latter depend-\ning on the direction of the spin current. (b) Schematics of\nthe sample geometry for XFMR. The sample (red disk) is\nmounted on the signal line (the central strip) of a coplanar\nwaveguide (CPW). The magnetization m(t) precesses about\nHe\u000b, driven by the in-plane continuous RF \feld h(t) in the\nCPW. The cone angle of precession is exaggerated for clar-\nity; its typical magnitude is \u00181\u000e. Circularly polarized X-ray\npulses from the synchrotron impinge at an grazing incidence\nangle on the sample in transverse geometry in order to enable\nstroboscopic detection of the oscillatory component of m(t)\nat variable phase delays.\ne\u000bect [57]. However, before performing the XFMR ex-\nperiment, the static magnetization of the samples has\nto be precharacterized with standard techniques, such as\nsuperconducting quantum interference device (SQUID)\nmagnetometry to measure the hysteresis loops along the\neasy and hard direction, followed by standard FMR mea-\nsurements.\nA. VNA-FMR\nVector network analyzers (VNA)-FMR measurements\nare used to characterize the magnetic resonances in order\nto judge whether these are suitable and intense enough\nfor the XFMR measurements at the synchrotron. VNA-\nFMR is a broadband FMR technique, where the sample is\nmounted onto a coplanar waveguide (CPW) and driven\nby an external RF \feld, while under a static magnetic\nbias \feld. Measurement of the S-parameters of the sam-\nple results in a frequency-\feld map, where the resonances\nappear in the form of Kittel curves (Eq. (2)). The angu-\nlar dependence of the resonances gives information about\nthe magnetic anisotropy [53]. It allows us to chose the\nbest azimuthal angle of the applied \feld with respect to\nthe crystallographic axes to separate the magnetic res-\nonances at the optimal distance for detecting STT [28].\nHence, at a given RF frequency, this gives us the cor-\nresponding \feld values of the resonances in the XFMR\nexperiments. Conventional FMR will normally probe the\nwhole thickness of a thin \flm since the skin depth of, e.g.,metallic iron at 10 GHz, is on the order a micron.\nB. XMCD\nAt the synchrotron, \frst the static XMCD is measured\nby sweeping the photon energy across the absorption\nedge of the magnetic elements. This allows us to se-\nlect the \fxed photon energies suitable for XFMR. The\nstatic XMCD is obtained from the di\u000berence between\ntwo X-ray absorption spectra recorded with the helic-\nity vector of the circularly polarized X-rays parallel and\nantiparallel, respectively, to the applied magnetic \feld\n[58]. The XMCD signal is proportional to the projection\nof the helicity vector, which is along the incident beam\ndirection ^k, onto the sample magnetization M, hence\nIXMCD/^k\u0001M.\nThe XMCD at the soft X-ray absorption edges, such\nas the Fe, Co, and Ni L2;3, is very strong [59], which\nhelps to compensate for the small changes in magneti-\nzation direction due to the limited cone angle ( <1\u000e) of\nthe precession in XFMR. The X-ray penetration length,\nwhich limits the sampling depth, is in the nm range, e.g.,\nfor pure Fe it is\u001820 nm at the Fe L3maximum at\u0018707\neV [57]. By tuning the photon energy away from the\nabsorption maximum the penetration length can be in-\ncreased (\u0018600 nm below the edge at 700 eV). Note that\nthe length scale of the probing depth is well matched to\nthe thickness of the magnetic layers in spin valves. The\ntypical lateral spot size of the X-ray beam on the sample\nis 200\u000220\u0016m2, again well suited for small devices.\nC. Time-resolved measurements\nThe measurement of the projected magnetic moment\nin XFMR does not require to take the di\u000berence between\nopposite circular polarizations as done in XMCD. In-\nstead, a change in the projection of the magnetization\nprecession is measured using a \fxed circular polarization.\nXFMR can be measured in two distinctly di\u000berent ge-\nometries, namely ( i) time-averaged in longitudinal ge-\nometry [12, 20] or ( ii) time-resolved in transverse geom-\netry [8, 25, 28]. In longitudinal geometry, a shortening\nof the magnetization vector along the z-axis (parallel to\nthe X-ray beam direction) leads to a di\u000berence \u0001 Mz=\nMs(1\u0000cos\u0012)\u00191\n2Ms\u00122, where\u0012is the small cone angle of\nthe magnetization precession. The time-averaged XFMR\nrequires no synchronization with the synchrotron clock,\ntherefore it can be done at an arbitrary frequency, but\nit needs a larger RF power which can lead to nonlinear\ne\u000bects.\nOnly measurements in transverse geometry give access\nto the precessional phase. This geometry is depicted in\nFig. 1(b). The transverse component of the magnetiza-\ntion precession will give a sinusoidal variation on top of\nthe static X-ray absorption signal. With the incident X-\nray beam perpendicular to the bias \feld, the oscillating4\ncomponent of the magnetization precession is measured\nwith a magnitude jMyj=Mssin\u0012\u0019Ms\u0012. Thus, for a\ntypical cone angle of \u0012\u00191\u000e, the transverse geometry\ngives a signal that is larger by a factor of \u0018200 com-\npared to the longitudinal geometry. Due to the shape\nanisotropy of the \flm, the precession is strongly ellipti-\ncal, often with a larger in-plane amplitude. This favors\na measurement geometry with the X-rays at grazing in-\ncidence. A good compromise is an X-ray incidence angle\nof\u001835\u000ewith respect to the plane of the sample, which\nensures that the signal is sensitive to the larger in-plane\ncomponent of the magnetization precession.\nUsing a vector magnet system, such as the portable oc-\ntupole magnet system (POMS) at Diamond [57], where\nthe \feld can be applied in any direction, permits a simple\nchange of the \feld from ( i) parallel to the photon direc-\ntion, as needed for static XMCD scans, to ( ii) orthogonal\nto both X-ray beam and RF \feld direction, as required\nfor time-resolved XFMR.\nThe detection of the X-ray absorption can be done by\neither X-ray transmission [27, 31], \ruorescence yield [23],\nor X-ray scattering or re\rectivity [2, 35, 36, 60, 61]. How-\never, RF plays havoc with total-electron yield. In the\ncase of transmission, the incident X-rays impinge on the\nsample through a hole in the signal line of the CPW.\nAfter passing through the sample, the transmitted X-\nrays are detected with X-ray excited optical luminescence\n(XEOL) emerging from the MgO or sapphire (Al 2O3)\nsubstrate using a photodiode placed behind the sample.\nNote that not all substrates, such as non-transparent ones\nlike Si, are suitable for XEOL detection [62].\nTime resolution is established by using the periodic\nX-ray pulses from the synchrotron (normally operating\nin multibunch or hybrid mode). To enable stroboscopic\nprobing, the RF driving \feld is taken as a harmonic of\nthe X-ray pulse frequency, hence the resonance is driven\nat multiples of the master oscillator clock of the stor-\nage ring. These harmonics are generated using an RF\ncomb generator (Atlantic Microwave) driven by the mas-\nter oscillator clock, which has a frequency of 499.65 MHz\n(at the DLS, ALS, and BESSY synchrotron). This cor-\nresponds to\u00182 ns intervals between consecutive X-ray\npulses, which have a pulse width of \u001835 ps (at DLS and\nBESSY) or\u001870 ps (at ALS). The desired frequency is\nselected using \flters and ampli\fers to drive a narrow\nband, high power (25{30 dBm) RF \feld to the CPW.\nA programmable delay line (Colby Instruments) enables\nphase shifting of the RF oscillation with respect to the\nX-ray pulses with a step resolution of \u00180.5 ps. De-\npending on the speci\fc technique either the transmitted,\ndi\u000bracted, or re\rected X-rays are measured using a pho-\ntodiode. Fig. 2 show a schematic representation of the\nsetup for DFMR; for XFMR and RFMR the electronics\nis very similar. The signal is obtained using a lock-in\nampli\fer (LIA) by switching the signal at a given audio\nfrequency. There are two usual modulation modes. In\namplitude modulation the LIA measures the di\u000berence\nbetween signals obtained with the RF signal on and o\u000b.\nFIG. 2. Schematic of the setup for DFMR measurements in\nthe RASOR di\u000bractometer at the Diamond Light Source. The\nsample is placed on the CPW, which is mounted on the cold\n\fnger inside the di\u000bractometer. Incident circularly or linearly\npolarized X-rays are scattered o\u000b the sample and detected via\na photodiode in a #-2#geometry. A variable magnetic \feld\nis applied in the scattering plane via a pair of permanent\nmagnets whose distance can be controlled externally. An RF\nsignal is fed to the CPW to drive the ferromagnetic resonance\nin the magnetic sample. As the synchrotron gives X-ray pulses\nat a frequency of \u0018500 MHz, a comb generator is used to\nproduce higher harmonics, which are selected and fed to the\nCPW. To probe the time dependence of the scattered X-ray\nintensity, a tunable delay line is used, which shifts the phase\nbetween the pump (the RF signal) and the probe (the pulsed\nX-rays). (Adapted from Ref. [63]).\nIn 180\u000e-phase modulation, the LIA measures the di\u000ber-\nence between signals obtained with the RF of opposite\nphase.\nD. XFMR\nIn order to record the time-resolved XAS signal with\ncircular polarization at \fxed photon energy, the RF fre-\nquency is locked to a multiple of the synchrotron clock.\nThen at \fxed angles and for given temperature, this\nleaves two free scanning parameters, namely the mag-\nnetic bias \feld strength and the delay time between X-ray\npulses and RF \feld.\nMagnetic \feld scans record the signal by sweeping the\nbias \feld at a constant delay time. The signal contains\nboth real and imaginary parts of the magnetic suscepti-\nbility, whose relative contributions strongly change across\nresonance. By measuring two \feld scans, which di\u000ber by5\n90\u000ein phase (obtained using the corresponding time de-\nlays), and \ftting these scans simultaneously using the\nKramers-Kronig relation, gives a good apprehension of\nthe \feld dependence of the resonances [40].\nDelay scans record the signal for each of the magnetic\nlayers at constant bias \feld by sweeping the delay time.\nAs an example, Fig. 3(a) shows a series delay scans over\ntwo periods of the phase taken at di\u000berent bias \felds (40{\n200 mT) across the Co resonance in a magnetic tunnel\njunction (MTJ), in more detail discussed in Sec. IV A.\nThe solid lines represent sinusoidal \fts to the experimen-\ntal data (dots), from which the amplitude and relative\nphase of the magnetization precession can be extracted.\nA sinusoidal function of the form\nS(t) =Xsin(2\u0019\u0017t) +Ycos(2\u0019\u0017t); (5)\nis \ftted to the delay scan, where tis the time delay and\n\u0017the frequency of the RF. This procedure is repeated\nfor various \feld strengths and directions. By extracting\nthe coe\u000ecients XandYin Eq. (5) from the delay scans,\nthe amplitude Aand phase of the oscillations can be\ndetermined using the relationships\nA=p\nX2+Y2; = 2 arctan\u0012Y\nA+X\u0013\n:(6)\nXFMR precessional plots are assembled by combining\nthe amplitudes and phases extracted from the delay scans\nmeasured over a range of bias \felds. This gives the \feld\ndependence of the amplitude and phase for each element\n(e.g., for Co and Ni in Fig. 3(b)), from which the type\nof coupling between layers can be assessed. By normaliz-\ning the XFMR signal to the static XMCD, the amplitude\nof the signal can be obtained per atom for each chemi-\ncal element in the sample. This enables a quantitative\ndecomposition of the resonance features [31].\nThe static coupling (i.e., exchange interaction) and dy-\nnamic coupling (i.e., spin pumping) give a very di\u000ber-\nent XFMR response, as can be understood from Eq. (4).\nConsider a pump layer FM1 that is free to rotate, and a\nprobe layer FM2 that is pinned. Using XFMR at a \fxed\nfrequency, we scan the \feld across the entire resonance.\nAt resonance, FM1 will show a symmetric peak for the\namplitude, while the phase is 90\u000edelayed with respect\nto the RF driving \feld. Across the entire resonance, the\nphase will change by 180\u000e. To investigate the type of\ncoupling between both layers we measure the XFMR re-\nsponse of FM2 at the resonance condition of FM1.\nFor static exchange coupling, E=\u0000Aexm1\u0001m2, so\nthatHe\u000b;2/m1. This means that the e\u000bective \feld in\nthe second layer is aligned along the magnetization of the\n\frst layer. Then the \feld dependent precession of FM2\nwill show a dispersive (bipolar) peak in the amplitude\nand a symmetric (unipolar) peak in the phase.\nOn the other hand, for dynamic exchange coupling\nHe\u000b;2/_ m2=\u0000i!m1. The magnetic \feld is imagi-\nnary, resulting in a 90\u000ephase change. In this case, the\n\feld dependent precession of FM2 will show a unipolar\npeak in the amplitude and bipolar peak in the phase.This behavior means that XFMR can distinguish be-\ntween static and dynamic coupling by their amplitude\nand phase signature in the precessional plot, and thus\ndetermine the relative contribution of these couplings.\nThis has previously been utilized for, e.g., exchange cou-\npled layers [4, 31], spin values [25], MgO magnetic tunnel\njunctions [29, 45], topological insulators [27, 30, 41], spin\nvalve with\u000e-layer [26], Heusler alloys [40], NiO antiferro-\nmagnetic interlayer [42], exchange springs [44], and \u000b-Sn\nthin \flms [47].\nE. DFMR and RFMR\nDFMR and RFMR measurements have been per-\nformed in the RASOR soft X-ray di\u000bractometer on beam-\nline I10 at the Diamond Light Source [57] (see setup in\nFig. 2). Incident X-rays with wavevector kiilluminate\nthe sample, while the scattered beam ( ks) is detected us-\ning a photodiode. The scattering geometry is con\fgured\nto probe the sample at certain di\u000braction or specular re-\n\rectivity conditions. The sample in the di\u000bractometer is\nmounted on a CPW that is connected to a liquid He cryo-\nstat arm which can reach temperatures down to 12 K. A\nbias magnetic \feld is applied by two permanent magnets,\nwhich can be positioned to vary both the \feld strength\nup to 200 mT and the orientation within the scatter-\ning plane. Perpendicular to the bias \feld, a transverse\nRF \feld around the central conductor of the CPW is\ngenerated, which excites the magnetization dynamics in\nthe system. In contrast to conventional XFMR measure-\nments, where the sample is mounted \rip-chip onto the\nCPW, in the scattering geometry the sample is mounted\nface up to allow for the X-ray beam to probe its sur-\nface. To ensure good coupling between the CPW and the\nprobed top surface, the sample must either be thinned,\nor in the case of multilayers, grown on a thin substrate\nof the order of 100 \u0016m.\nIn DFMR, where the detector is aligned to a Bragg\npeak or magnetic scattering peak, the stroboscopic sig-\nnal is used to measure delay scans for di\u000berent linear or\ncircular polarization of the incident X-rays, to give infor-\nmation about the periodic spin structure.\nIn RFMR where the photo diode detector accepts the\nre\rected beam the stroboscopic signal is used to measure\ndelay scans for di\u000berent values of the scattering vector\nQz, to obtain depth information. An advantage of RFMR\nover DFMR is that it can be done on thin \flms and\nmultilayers, and no single crystals are needed.\nIV. X-RAY BASED FMR EXAMPLES\nA. XFMR of spin-current mediated exchange\ncoupling in MgO-based MTJs\nMagnetic tunnel junctions composed of ferromagnetic\nlayers which are mutually interacting through a nonmag-6\nnetic spacer layer are at the core of magnetic sensor and\nmemory devices. G ladczuk et al. [45] used layer-resolved\nXFMR to investigate the coupling between the magnetic\nlayers of a Co/MgO/Py MTJ. Two magnetic resonance\npeaks were observed for both magnetic layers, as probed\nat the Co and Ni L3X-ray absorption edges.\nFigure 3 shows XFMR delay scans for the Co layer in\nthe Co/MgO/Py MTJ at 80 K continuously driven at 4\nGHz. The curves in Fig. 3(a) show a strong increase in\namplitude as well as a large phase shift across the reso-\nnance at\u001890 mT. The amplitude and phase of the pre-\ncession, which are extracted using Eq. (6), are shown in\nFig. 3(b) and (c), respectively, for both Co (orange) and\nNi (blue) as a function of the bias \feld. The amplitude\ncurves show that the Ni resonance originating from the\nPy layer around \u0018120 mT is strongly coupled with the\nCo layer. On the other hand, the Co resonance around\n\u001890 mT is only weakly present in the Py layer. Instead\nof plotting amplitude Aand phase , one can also plot\nthe FMR signal in the ( X,Y)-plane as a function of \feld\n[45]. Since the sine and cosine functions in Eq. (5) are\northogonal, the estimators of XandYare given by pro-\njections to orthogonal subspaces.\nA theoretical model based on the Landau-Lifshitz-Gil-\nbert-Slonczewski equation (Eq. (4)) was developed, in-\ncluding exchange coupling and spin pumping between the\nmagnetic layers. Fits to the experimental data were car-\nried out, both with and without a spin pumping term,\nand the goodness of the \ft was compared using a likeli-\nhood ratio test. This rigorous statistical approach pro-\nvided an unambiguous proof of the existence of interlayer\ncoupling mediated by spin pumping through MgO [45].\nIt was also found that spin pumping is more e\u000bective\nat lower temperatures, which agrees with the theoretical\nunderstanding.\nB. XFMR of coherent spin currents in\nantiferromagnetic NiO\nAntiferromagnets have recently gained large interest in\nthe \feld of spintronics, as they allow for faster and more\nrobust memory operation than present technologies and\nas they can carry spin current over long distances. How-\never, many fundamental physics questions about these\nmaterials regarding their spin transport properties still\nremain unanswered [64]. A spin current generated by\nspin pumping should have a single wave mode, carrying\nthe coherent magnetization excitation. In contrast, spin\ncurrents generated by thermal gradients produce incoher-\nent currents with a continuum of spin excitation modes.\nThe magnetic excitations in antiferromagnets typically\nhave THz frequencies, while the resonant excitation of\nthe ferromagnetic injector is in the GHz range. Con-\nventional spin pumping experiments measure only the\ntime-averaged DC component of the spin current, i.e.,\nthey cannot distinguish between GHz and THz frequen-\ncies, which is needed to determine how the spin current\nFIG. 3. Time resolved precession. (a) Series of XFMR de-\nlay scans for the Co layer in a Co/MgO/Py MTJ at 80 K\ncontinuously driven at 4 GHz. As expected, the period of\nthe precession is 250 ps, and the delay scan covers two peri-\nods. For clarity, the data points (circles) obtained at di\u000berent\nmagnetic \feld values (between 40 and 200 mT) are shifted\nby a constant o\u000bset and have been di\u000berently colored. The\ndrawn lines represent the \ftted sinusoidal functions. Their\namplitude and phase as a function of magnetic \feld strength\nis plotted in panels (b) and (c), respectively, for both the Co\n(orange) and Py (blue) layers. (Adapted from G ladczuk et\nal.[45]).\npropagates. Alternative techniques such as XFMR are\nneeded to measure the time-varying AC spin current.\nDabrowski et al. [42] used XFMR to study the coher-\nent spin current propagation in a device with three layers,\nwhere the top (injector) and bottom (sink) layers were\nferromagnetic NiFe and FeCo, respectively, and the mid-\ndle layer was epitaxial NiO (001). The phase and am-7\nFIG. 4. DFMR delay scans of the structural and magnetic\npeaks as a function of linear polarization angle. Measure-\nments of (a,b) the anisotropic mode B at 6 GHz and (c,d)\nthe isotropic mode A at 2 GHz. The results for the magnetic\npeak and the structural (0,0,3) peak are shown in the left\nand the right column, respectively. The magnetic resonance\nmodes are probed with linearly polarized light for the range\nof incident polarization angles \u0011between 0{180\u000e. (Adapted\nfrom Ref. [60]).\nplitude of the magnetization precession within adjoining\nsource and sink FM layers were detected, from which\nthe injection and transmission of pure AC spin current\nthrough NiO can be inferred. It was found that magne-\ntization modes in the FM layers oscillate in phase. Fur-\nthermore, the e\u000eciency of the spin transfer varied with\nthe thickness of the antiferromagnet, with a maximal ef-\n\fciency for a 2-nm-thick layer. These results indicate\nthat a spin current propagates coherently through the\nantiferromagnetic NiO layer. The AC spin current is en-\nhanced for NiO thicknesses of less than 6 nm, both with\nand without a nonmagnetic spacer layer inserted into the\nstack, in a manner consistent with previously reported\nexperimental measurements of DC spin current and the-\noretical studies [65]. The XFMR results show that the\npropagation of spin current through NiO layers is medi-\nated by evanescent antiferromagnetic spin wave modes at\nGHz frequencies, rather than THz frequency magnons.\nC. DFMR for mode-resolved detection of\nmagnetization dynamics\nRecent scienti\fc interest has shifted towards more\ncomplex magnetically ordered materials, which are\npromising for high-density and low-energy consumption\ndevices. These systems contain chiral magnetic phases\nsuch as helical, conical, or skyrmion spin structures,\noriginating from the Dzyaloshinskii-Moriya interaction(DMI) found in noncentrosymmetric bulk materials, as\nwell as in systems where symmetry breaking occurs at\na ferromagnetic/heavy metal interface. Such spin struc-\ntures are much more complex than simple ferromagnetic\nstructures, especially their dynamic behavior is so far ill-\nunderstood.\nThe periodic structure of magnetically ordered sys-\ntems can be probed by resonant elastic X-ray scattering\n(REXS), making use of interference e\u000bects from the reg-\nularly repeating magnetization density variations. This\nleads to pure magnetic X-ray scattering peaks which give\ninformation about the static magnetic structure. Analy-\nsis of these magnetic peaks in REXS measurements using\nsynchrotron radiation has led to signi\fcant progress in\nthe understanding of chiral magnetic systems [66, 67].\nIn a pioneering DFMR experiment, Burn et al. [60]\ninvestigated the complex dynamic behavior of the chi-\nral spin structure in Y-type hexaferrite Ba 2Mg2Fe12O22.\nVNA-FMR measurements of this material showed a \feld-\nfrequency map containing two ferromagnetic resonance\nmodes. While mode A is isotropic, i.e., its \feld value is\nindependent of the direction of the applied \feld, mode B\nis anisotropic, showing greater absorption at increasingly\nhigher \felds as the \feld direction rotates out-of-plane.\nREXS at the Fe L2;3absorption edge was used to char-\nacterize the static magnetic structure of the hexaferrite\nand to determine its \feld dependence. Static REXS mea-\nsurements along (0,0, `) in zero \feld show a (0,0,3) struc-\ntural peak decorated with two incommensurate magnetic\nsatellites.\nThe DFMR signal was measured by pointing the pho-\ntodiode at the scattered beam, selecting either the struc-\ntural or the magnetic satellite peak (Fig. 2). Delay scans\nwere measured as a function of applied \feld using linearly\npolarized X-rays. Sinusoidal \fts to the measured data en-\nables the extraction of amplitude and phase. Fig. 4 shows\nthe delay scans of the structural and magnetic peaks of\nthe Y-type hexaferrite for variable incident linear polar-\nization angles \u0011. The panels (a,b) in the top row refer to\nthe anisotropic mode B at 6 GHz, and the panels (c,d)\nin the bottom to the isotropic mode A at 2 GHz. The\nleft and right column refer to the results for the mag-\nnetic peak and the structural (0,0,3) peak, respectively.\nThe results were compared to computer simulations of\nthe Y-type hexaferrite to obtain insight in the periodic\nspin structure of this material.\nA second example of the use of DFMR for mode re-\nsolved detection concerns the dynamic behavior of topo-\nlogical spin textures and chiral magnets, which is an area\nof signi\fcant interest and key to the development of fast\nand e\u000ecient spintronics devices. DFMR measurements\nby Burn et al. [63] revealed how the time-dependence\nof the magnetization dynamics relate to the complex\nspin texture in the well-known chiral magnetic system\nCu2OSeO 3. Using polarized soft X-rays, the dynamic\nexcitations in all three dimensions were probed, which\nrevealed phase shifts that were previously undetectable\nand indistinguishable using conventional FMR.8\nFIG. 5. (a) Pseudo-3D plot of the RFMR signal and its pro-\njection showing the dynamic contribution to the re\rectivity\nfor a [CoFeB/MgO/Ta] 4multilayer as a function of pump-\nprobe time delay. The measurements were carried out with\nleft-circularly polarized X-rays at the Fe L3resonance (707.7\neV) and in an out-of-plane \feld of 29 mT using RF excitation\nat 2 GHz. The various delay curves are shown for di\u000berent Qz,\nranging between 0 and 0.6 nm\u00001. The color scale represents\nthe normalized intensity for each delay scan, highlighting the\nsinusoidal dependence and the shift in phase as Qzis varied\nwhen the intensity is small. (b) Static and dynamic re\rectiv-\nity, and (c) phase of the dynamic re\rectivity as a function of\nQz. The phase point size is scaled by the strength of the dy-\nnamic signal amplitude, and the blue-shaded regions indicate\nwhere the 180\u000ephase shifts have been subtracted to reveal the\notherwise smooth phase variation. (Adapted from Ref. [61]).\nD. RFMR on a [CoFeB/MgO/Ta] 4multilayer\nX-ray re\rectivity with the photon energy tuned to the\nabsorption edge has become a valuable tool for character-\nizing the depth-dependent structure of layered materials.\nThe X-ray re\rectivity is measured as a function of the\nscattering vector Qz=ks\u0000ki= (4\u0019=\u0015) sin#, where ki\n(ks) is the ingoing (outgoing) wavevector of the X-rays\nwith incident angle #and wavelength \u0015. The scattering\nlength density, which gives the scattering strength of the\nchemical and magnetic species within the depth pro\fle of\nthe \flm, is obtained through \ftting the re\rectivity data.\nBurn et al. [61] revealed the depth dependence of the\nmagnetization dynamics in a [CoFeB/MgO/Ta] 4multi-layer system. The structural depth pro\fle was charac-\nterized through static X-ray re\rectometry. The dynamic\nre\rectivity was probed with stroboscopic DFMR using\nan out-of-plane saturating \feld of HBias= 29 mT and\nan RF \feld generated by the CPW beneath the sample.\nThe RF \feld was phase-locked to the fourth harmonic of\nthe\u0018500 MHz synchrotron master clock at 2 GHz. The\ntime dependence of the re\rectivity during precession was\nmapped out as a function of the time delay between the\nRF pump and X-ray probe. Fig. 5(a) shows a color map\nof the sinusoidal variation in the re\rected signal with a\n500 ps period, corresponding to the 2 GHz excitation.\nThe amplitude and phase of the dynamic signal are ex-\ntracted by \ftting the sinusoidal delay scans. The ampli-\ntude is plotted in Fig. 5(b) alongside the static re\rectivity\nfor the di\u000berent values of Qz, ranging between 0 and 0.6\nnm\u00001, and the phase in Fig. 5(c).\nBoth the static intensity and the amplitude of the\ndynamic signal in Fig. 5(b) show re\rectivity fringes re-\nsulting from interference e\u000bects arising from the layered\nchemical and magnetic structure. Additional minima are\nobserved in the dynamic case. The phase of the dynamic\nsignal in Fig. 5(c) shows variations with two contribu-\ntions. Firstly, abrupt 180\u000ephase jumps occur, coincid-\ning with minima in the amplitude of the dynamic signal.\nThese 180\u000ejumps correspond to inversion of the sign of\nthe XMCD signal measured at di\u000berent scattering con-\nditions. In addition, there are smoother variations in\nthe phase, which can be attributed to variations in the\nmagnetization dynamics occurring between the magnetic\nlayers in the multilayered structure.\nTo reveal the depth dependent magnetization dynam-\nics, the experimental results were compared with model-\ning of the dynamic behavior In all layers, the magnetiza-\ntion precesses about a nominal static state when excited\nby an RF \feld. It was shown that inclusion of a small,\nbut signi\fcant phase lag of 5\u000ebetween the four layers\nis necessary to explain the observed change in phase of\nthe dynamic signal. In contrast, a single slab of mag-\nnetic thin \flm material shows a coherent precession of\nthe magnetization as a function of depth.\nWith RFMR, the dynamics from di\u000berent layers con-\ntaining the same element can be explored, and this tech-\nnique has the potential to study the dynamics of inter-\nfacial layers and proximity e\u000bects in complex thin \flm\nand multilayer materials for future magnetic memory and\nprocessing device applications.\nV. CONCLUSIONS\nAlthough conventional FMR is a powerful technique\nto study magnetic resonances in thin \flms and multi-\nlayers, the measured response corresponds to an aver-\nage over the entire magnetic structure of the sample. In\ncontrast, X-ray based FMR techniques allow for time-\nresolved measurements of the magnetization dynamics,\nand, in addition, o\u000ber the bene\fts of XMCD, such as9\nelement-, site-, and shell-speci\fcity [57]. The time reso-\nlution is achieved by stroboscopic probing using higher\nharmonics (1-10 GHz) of the synchrotron master clock.\nXFMR can be used to study spin-transfer torque, dipo-\nlar \feld strength, magneto-crystalline anisotropy, inter-\nlayer exchange coupling, gyromagnetic ratio and damp-\ning constants. It can be applied to study the behav-\nior of spintronics systems, e.g., spin pumping in mag-\nnetic multilayers, heterostructures, spin valves, MTJ, etc.\nThe amplitude and phase of the magnetic resonances\nextracted from the \feld-dependence of the precessional\nplots enable us to distinguish between static and dy-\nnamic exchange coupling and to quantify their relative\ncontributions. Apart from measuring the signal in ab-\nsorption, XFMR can also be detected in di\u000braction and\nre\rectivity; each of these techniques bringing unique ad-vantages. DFMR reveals the dynamical spin modes at\nthe probed magnetic wavevectors, and RFMR gives the\ndepth-resolved dynamics in magnetic multilayers. Future\nXFMR studies can be envisaged to investigate vortex dy-\nnamics, spatial resolution imaging, and X-ray hologra-\nphy.\nVI. ACKNOWLEDGMENTS\nThe XFMR experiments were carried out on beamline\nI10 at the Diamond Light Source (Oxfordshire, United\nKingdom). We like to acknowledge valuable collabo-\nrations with Alex A. Baker, David M. Burn, Maciej\nDabrowski, Adriana I. Figueroa, Lukasz Gladczuk, and\nRobert J. Hicken.\n||||||||{\n[1] G. van der Laan, J. Electron Spectrosc. Relat. Phenom.\n220 (2017) 137{146.\n[2] D. M. Burn, S. L. Zhang, G. van der Laan, T. Hesjedal,\nAIP Advances 11 (2021) 015327.\n[3] W. E. Bailey, L. Cheng, D. J. Keavney, C.-C. Kao,\nE. Vescovo, D. A. Arena, Phys. Rev. B 70 (2004) 172403.\n[4] D. A. Arena, E. Vescovo, C. C. Kao, Y. Guan, W. E.\nBailey, Phys. Rev. B 74 (2006) 064409.\n[5] Y. Guan, W. E. Bailey, C.-C. Kao, E. Vescovo, D. A.\nArena., J. Appl. Phys. 99 (2006) 08J305.\n[6] D. A. Arena, E. Vescovo, C.-C. Kao, Y. Guan, W. E.\nBailey, J. Appl. Phys. 101 (2007) 09C109.\n[7] Y. Guan, W. E. Bailey, E. Vescovo, C.-C. Kao, D. A.\nArena, J. Magn. Magn. Mater. 312 (2007) 437{378.\n[8] D. A. Arena, Y. Ding, E. Vescovo, S. Zohar, Y. Guan,\nW. E. Bailey, Rev. Sci. Instrum. 80 (2009) 083903.\n[9] W. Bailey, C. Cheng, R. Knut, O. Karis, S. Au\u000bret, S. Zo-\nhar, D. Keavney, P. Warnicke, J.-S. Lee, D. A. Arena,\nNat. Commun. 4 (2013) 2025.\n[10] P. Warnicke, R. Knut, E. Wahlstr om, O. Karis, W. E.\nBailey, D. A. Arena, J. Appl. Phys. 113 (2013) 033904.\n[11] P. Warnicke, E. Stavitski, J.-S. Lee, A. Yang, Z. Chen,\nX. Zuo, S. Zohar, W. E. Bailey, V. G. Harris, D. A.\nArena, Phys. Rev. B 92 (2015) 104402.\n[12] J. Goulon, A. Rogalev, F. Wilhelm, N. Jaouen,\nC. Goulon-Ginet, G. Goujon, J. Ben Youssef, M. V. In-\ndendom, JETP Lett. 82 (2005) 696{701.\n[13] J. Goulon, A. Rogalev, F. Wilhelm, N. Jaouen,\nC. Goulon-Ginet, C. Brouder, Eur. Phys. J. B 53 (2006)\n169{184.\n[14] J. Goulon, A. Rogalev, F. Wilhelm, C. Goulon-Ginet,\nG. Goujon, J. Synchrotron Rad. 14 (2007) 257{271.\n[15] A. Rogalev, J. Goulon, F. Wilhelm, C. Brouder,\nA. Yaresko, J. Ben Youssef, M. V. Indenbom, J. Magn.\nMagn. Mater. 321 (2009) 3945{3962.\n[16] J. Goulon, A. Rogalev, F. Wilhelm, G. Goujon,\nC. Brouder, A. Yaresko, J. Ben Youssef, M. V. Inden-\nbom, J. Magn. Magn. Mater. 322 (2010) 2308{2329.\n[17] J. Goulon, A. Rogalev, G. Goujon, F. Wilhelm, J. Ben\nYoussef, C. Gros, J.-M. Barbe, R. Guilard, Int. J. Mol.\nSci. 12 (2011) 8797{8835.[18] J. Goulon, C. Brouder, A. Rogalev, G. Goujon, F. Wil-\nhelm, J. Magn. Magn. Mater. 366 (2014) 1{23.\n[19] G. Boero, S. Rusponi, P. Bencok, R. S. Popovic,\nH. Brune, P. Gambardella, Appl. Phys. Lett. 87 (2005)\n152503.\n[20] G. Boero, S. Mouaziz, S. Rusponi, P. Bencok, F. Nolting,\nS. Stepanow, P. Gambardella, New J. Phys. 10 (2008)\n013011.\n[21] G. Boero, S. Rusponi, P. Bencok, R. Meckenstock, J.-\nM. Thiele, F. Nolting, P. Gambardella, Phys. Rev. B 79\n(2009) 224425.\n[22] G. Boero, S. Rusponi, J. Kavich, A. Lodi Rizzini, C. Pi-\namonteze, F. Nolting, C. Tieg, J.-U. Thiele, P. Gam-\nbardella, Rev. Sci. Instrum. 80 (2009) 123902.\n[23] M. K. Marcham, P. S. Keatley, A. Neudert, R. J. Hicken,\nS. A. Cavill, L. R. Shelford, G. van der Laan, N. D.\nTelling, J. R. Childress, J. A. Katine, P. Shafer, E. Aren-\nholz, J. Appl. Phys. 109 (2011) 07D353.\n[24] M. K. Marcham, W. Yu, P. S. Keatley, L. R. Shelford,\nP. Shafer, S. A. Cavill, H. Qing, A. Neudert, J. R. Chil-\ndress, J. A. Katine, E. Arenholz, N. D. Telling, G. van der\nLaan, , R. J. Hicken, Appl. Phys. Lett. 102 (2013) 062418.\n[25] M. K. Marcham, L. R. Shelford, S. A. Cavill, P. S. Keat-\nley, W. Yu, P. Shafer, A. Neudert, J. R. Childress, J. A.\nKatine, E. Arenholz, N. D. Telling, G. van der Laan, R. J.\nHicken, Phys. Rev. B 87 (2013) 180403(R).\n[26] J. Li, L. R. Shelford, P. Shafer, A. Tan, J. X. Deng, P. S.\nKeatley, C. Hwang, E. Arenholz, G. van der Laan, R. J.\nHicken, Z. Q. Qiu, Phys. Rev. Lett. 117 (2016) 076602.\n[27] A. A. Baker, A. I. Figueroa, L. J. Collins-McIntyre,\nG. van der Laan, T. Hesjedal, Sci. Rep. 5 (2015) 7907.\n[28] A. A. Baker, A. I. Figueroa, C. J. Love, S. A. Cavill,\nT. Hesjedal, G. van der Laan, Phys. Rev. Lett. 116 (2016)\n047201.\n[29] A. A. Baker, A. I. Figueroa, D. Pingstone, V. K. Lazarov,\nG. van der Laan, T. Hesjedal, Sci. Rep. 6 (2016) 35582.\n[30] A. I. Figueroa, A. A. Baker, L. J. Collins-McIntyre,\nT. Hesjedal, G. van der Laan, J. Magn. Magn. Mater.\n400 (2016) 178.\n[31] G. B. G. Stenning, L. R. Shelford, S. A. Cavill, F. Ho\u000b-\nmann, M. Haertinger, T. Hesjedal, G. Woltersdorf, G. J.10\nBowden, S. A. Gregory, C. H. Back, P. A. J. de Groot,\nG. van der Laan, New J. Phys. 17 (2015) 013019.\n[32] P. Klaer, F. Ho\u000bmann, G. Woltersdorf, E. Arbelo Jorge,\nM. Jourdan, C. H. Back, H. J. Elmers, J. Phys. D: Appl.\nPhys. 44 (2011) 425004.\n[33] T. Martin, G. Woltersdorf, C. Stamm, H. A. D urr,\nR. Mattheis, C. H. Back, G. Bayreuther, J. Appl. Phys.\n103 (2008) 07B112.\n[34] T. Martin, G. Woltersdorf, C. Stamm, H. A. D urr,\nR. Mattheis, C. H. Back, G. Bayreuther, J. Appl. Phys.\n105 (2009) 07D310.\n[35] R. Salikhov, R. Abrudan, F. Br ussing, S. Buschhorn,\nM. Ewerlin, D. Mishra, F. Radu, I. A. Garifullin,\nH. Zabel, Appl. Phys. Lett. 99 (2011) 092509.\n[36] R. Salikhov, R. Abrudan, F. Br ussing, K. Gross, C. Luo,\nK. Westerholt, H. Zabel, F. Radu, I. A. Garifullin, Phys.\nRev. B 86 (2012) 144422.\n[37] K. W. Chou, A. Puzic, H. Stoll, G. Sch utz, B. Van\nWaeyenberge, T. Tyliszczak, K. Rott, G. Reiss,\nH. Br uckl, I. Neudecker, D. Weiss, C. H. Back, J. Appl.\nPhys. 99 (2006) 08F305.\n[38] K. Ollefs, R. Meckenstock, D. Spoddig, F. M. R omer,\nC. Hassel, C. Sch oppner, V. Ney, M. Farle, A. Ney, J.\nAppl. Phys. 117 (2015) 223906.\n[39] S. Bonetti, R. Kukreja, Z. Chen, D.Spoddig, K. Ollefs,\nC. Sch oppner, R. Meckenstock, A. Ney, J. Pinto,\nR. Houanche, J. Frisch, J. St ohr, H. A. D urr, H. Ohldag,\nRev. Sci. Instrum. 86 (2015) 093703.\n[40] C. J. Durrant, L. R. Shelford, R. A. J. Valkass, R. J.\nHicken, A. I. Figueroa, A. A. Baker, L. Du\u000by, G. van\nder Laan, P. Shafer, E. Arenholz, C. Klewe, S. A. Cavill,\nJ. R. Childress, J. A. Katine, Phys. Rev. B 96 (2017)\n144421.\n[41] A. A. Baker, A. I. Figueroa, T. Hesjedal, G. van der Laan,\nJ. Magn. Magn. Mater. 473 (2019) 470{476.\n[42] M. Dabrowski, T. Nakano, D. Burn, A. Frisk, D. G. New-\nman, C. Klewe, Q. Li, M. Yang, P. Shafer, E. Arenholz,\nT. Hesjedal, G. van der Laan, Z. Q. Qiu, R. J. Hicken,\nPhys. Rev. Lett. 124 (2020) 217201.\n[43] C. Klewe, Q. Li, M. Yang, A. T. N'Diaye, D. M. Burn,\nT. Hesjedal, A. I. Figueroa, C. Hwang, J. Li, R. J. Hicken,\nP. Shafer, E. Arenholz, G. van der Laan, Z. Qiu, Syn-\nchrotron Radiat. News 33 (2020) 12{19.\n[44] M. Dabrowski, A. Frisk, D. M. Burn, D. G. Newman,\nC. Klewe, A. T. N'Diaye, P. Shafer, G. J. Bowden,\nT. Hesjedal, G. van der Laan, G. Hrkac, R. J. Hicken,\nACS Appl. Mater. Interfaces 12 (2020) 52116{52124.\n[45] L. Gladczuk, L. Gladczuk, P. Dluzewski, K. Lasek,\nP. Aleshkevych, D. M. Burn, G. van der Laan, T. Hes-\njedal, Phys. Rev. B 103 (2021) 064416.[46] M. Dabrowski, R. J. Hicken, A. Frisk, D. G. Newman,\nC. Klewe, A. N'Diaye, P. Shafer, G. van der Laan, T. Hes-\njedal, G. J. Bowden, New J. Phys. 23 (2021) 023017.\n[47] L. Gladczuk, L. Gladczuk, P. Dluzewski, G. van der\nLaan, T. Hesjedal, Phys. Status Solidi RRL 15 (2021)\n2100137.\n[48] X. Yang, J.-F. Cao, J.-Q. Li, F.-Y. Zhu, R. Yu, J. He, Z.-\nL. Zhao, Y. Wang, R.-Z. Tai, Nucl. Sci. Tech. 33 (2022)\n63.\n[49] C. Klewe, S. Emori, Q. Li, M. Yang, B. A. Gray, H.-\nM. Jeon, B. M. Howe, Y. Suzuki, Z. Q. Qiu, P. Shafer,\nE. Arenholz, New J. Phys. 24 (2022) 013030.\n[50] Y. Lim, S. Wu, D. A. Smith, C. Klewe, P. Shafer, S. Emo,\narXiv 2208.07294 (2022).\n[51] S. Emori, C. Klewe, J.-M. Schmalhorst, J. Krieft,\nP. Shafer, Y. Lim, D. A. Smith, A. Sapkota, A. Srivas-\ntava, C. Mewes, Z. Jiang, B. Khodadadi, H. Elmkharram,\nJ. J. Heremans, E. Arenholz, G. Reiss, T. Mewes, Nano\nLett. 20 (2020) 7828{7834.\n[52] Y. Pogoryelov, M. Pereiro, S. Jana, A. Kumar,\nS. Akansel, M. Ranjbar, D. Thonig, D. Primetzhofer,\nP. Svedlindh, J. Akerman, O. Eriksson, O. Karis, D. A.\nArena, Phys. Rev. B 101 (2020) 054401.\n[53] A. H. Morrish, The Physical Principles of Magnetism,\nJohn Wiley & Sons, New York, 1965.\n[54] J. C. Slonczewski, J. Magn. Magn. Mater 159 (1996) L1{\nL7.\n[55] L. Berger, Phys. Rev. B 54 (1996) 9353{9358.\n[56] Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, B. I.\nHalperin, Rev. Mod. Phys. 77 (2005) 1375.\n[57] G. van der Laan, A. I. Figueroa, Coord. Chem. Rev. 277-\n278 (2014) 95{129.\n[58] G. van der Laan, J. Phys.: Conf. Ser. 430 (2013) 012127.\n[59] G. van der Laan, B. T. Thole., Phys. Rev. B 43 (1991)\n13401{13411.\n[60] D. M. Burn, S. Zhang, K. Zhai, Y. Chai, Y. Sun, G. van\nder Laan, T. Hesjedal, Nano Lett. 20 (2020) 345{352.\n[61] D. M. Burn, S. L. Zhang, G. Q. Yu, Y. Guang, H. J.\nChen, X. P. Qiu, G. van der Laan, T. Hesjedal, Phys.\nRev. Lett. 125 (2020) 137201.\n[62] C. A. F. Vaz, C. Mouta\fs, M. Buzzi, J. Raabe, J. Elec-\ntron Spectrosc. Relat. Phenom. 189 (2013) 1{4.\n[63] D. M. Burn, S. L. Zhang, G. van der Laan, T. Hesjedal,\nPhys. Rev. B 106 (2022) 174409.\n[64] H. Reichlova, R. Schlitz, S. T. B. Goennenwein, Physics\n13 (2020) 83.\n[65] R. Khymyn, I. Lisenkov, V. S. Tiberkevich, A. N. Slavin,\nB. A. Ivanov, Phys. Rev. B 93 (2016) 224421.\n[66] G. van der Laan, C. R. Physique 9 (2008) 570{584.\n[67] S. L. Zhang, A. Bauer, H. Berger, C. P\reiderer, G. van\nder Laan, T. Hesjedal, Phys. Rev. B 93 (2016) 214420." }, { "title": "1512.01392v1.Itinerant_ferromagnetism_in_1D_two_component_Fermi_gases.pdf", "content": "Itinerant ferromagnetism in 1D two-component Fermi gases\nYuzhu Jiang,1D.V. Kurlov,2Xi-Wen Guan,1, 3F. Schreck,2and G.V. Shlyapnikov1, 2, 4, 5\n1State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics,\nWuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, Wuhan 430071, China\n2Van der Waals-Zeeman Institute, Institute of Physics, University of Amsterdam,\nScience Park 904, 1098 XH Amsterdam, The Netherlands\n3Department of Theoretical Physics, Research School of Physics and Engineering,\nAustralian National University, Canberra ACT 0200, Australia\n4LPTMS, CNRS, Univ. Paris-Sud, Universit\u0013 e Paris-Saclay, 91405 Orsay, France\n5Russian Quantum Center, Novaya Street 100, Skolkovo, Moscow Region 143025, Russia\nWe study a one-dimensional two-component atomic Fermi gas with an in\fnite intercomponent\ncontact repulsion. It is found that adding an attractive resonant odd-wave interaction breaking\nthe rotational symmetry one can make the ground state ferromagnetic. A promising system for\nthe observation of this itinerant ferromagnetic state is a 1D gas of40K atoms, where 3D s-wave\nandp-wave Feshbach resonances are very close to each other and the 1D con\fnement signi\fcantly\nreduces the inelastic decay.\nItinerant ferromagnetism of degenerate spin-1/2\nfermions is an intriguing problem promoting our under-\nstanding of strongly correlated systems [1]. The origin\nof such ferromagnetic states is deeply rooted in quantum\nmechanics. In contrast to ultracold bosons, degenerate\nfermions try to avoid the ferromagnetic state because it\nrequires them to have a signi\fcantly higher kinetic en-\nergy than in non-ferromagnetic states. Ultracold gases\nof atomic fermions which are in two internal states can\nbe mapped onto spin-1/2 fermions treating the internal\nenergy levels as pseudo-spin states. The ferromagnetic\nphase is the one where all atoms are in the same super-\nposition of the two internal states, and one has a system\nof identical fermions. The kinetic energy is then higher\nthan, for example, in the paramagnetic phase represent-\ning a statistical mixture of the two spin components.\nItinerant ferromagnetism for fermions is studied since\nthe 1930-ths, when the Stoner criterion for ferromag-\nnetism in a free electron gas was introduced [2, 3]. In\nthree dimensions the ground state can be ferromagnetic\nif there is a strong intercomponent repulsion in the para-\nmagnetic state, which compensates the large di\u000berence in\nthe kinetic energies of the ferro- and paramagnetic states.\nFor atomic Fermi gases the Stoner mechanism was dis-\ncussed in a number of papers [4] and was then tested by\nMonte Carlo calculations [5, 6] which found an instability\non approach to the strongly interacting regime. The ef-\nforts to stabilize the ferromagnetic state experimentally\ndid not succeed [7, 8]. The reason is [7{9] that in three\ndimensions a large intercomponent repulsion corresponds\nto a very large and positive s-wave scattering length, and\nthere is a weakly bound dimer of two fermions belonging\nto di\u000berent internal states. A composition of such dimers\nhas a lower energy than the ferromagnetic state even if\nthe dimers overlap with each other, and at common gas\ndensities the formation of the dimerized phase out of the\ncloud of atoms is very fast.\nIn one and two dimensions the di\u000berence in the ki-netic energies of the ferro- and non-ferromagnetic states\nis even larger than in 3D. For example, in the 1D two-\ncomponent Fermi gas even at an in\fnite intercomponent\ncontact repulsion the energies of these states are equal,\nand at any \fnite repulsion the antiferro- or paramag-\nnetic states have a lower energy [10]. Therefore, it looks\nsuch that in low dimensions making the ground state\nferromagnetic is harder than in 3D. However, we show\nthat in 1D the odd-wave interaction (analog of p-wave in\nhigher dimensions) can drastically change the situation\nand make the ground state ferromagnetic. This interac-\ntion is momentum dependent and does not fall into the\nclass of interactions satisfying the conditions of the Lieb-\nMattis theorem [11] which states that in the 1D (two-\ncomponent) Fermi system the ground state can not be\nferromagnetic.\nTo make the odd-wave interaction signi\fcant one needs\na Feshbach and/or con\fnement-induced odd-wave reso-\nnance. At the same time, another resonance is needed in\norder to make the even-wave contact interaction strongly\nrepulsive. Here we have a \\present from nature\". In\nthe case of40K atoms the s-wave resonance for the in-\nteraction between 9 =2;\u00007=2 and 9=2;\u00009=2 states occurs\nat the magnetic \feld of 202.1 Gauss, and is very close\nto thep-wave resonance for the interaction between two\n9=2;\u00007=2 atoms at 198.8 Gauss [12, 13]. In the \felds\nbetween 198.8 Gauss and 202.1 Gauss the s-wave inter-\naction is repulsive and the p-wave interaction is attrac-\ntive. Moreover, the reduction of dimensionality to 1D\ndecreases the inelastic decay even not far from the res-\nonances, which is very promising for achieving itinerant\nferromagnetism in the 1D gas of40K.\nThe even-wave scattering of identical fermions requires\nthat the spatial wavefunction of colliding atoms is sym-\nmetric, and the spinor part is antisymmetric, i.e. it oc-\ncurs through the singlet channel. On the contrary, in\nthe odd-wave scattering the spatial part of the fermionic\nwavefunction is antisymmetric, and the spinor part isarXiv:1512.01392v1 [cond-mat.quant-gas] 4 Dec 20152\nsymmetric, which corresponds to the triplet state. If the\nodd-wave interaction is the same in all triplet states, and\nif it is relatively weak (such that in the expression for the\ncorresponding interaction amplitude we can keep only the\nlowest term, which is proportional to k2, withkbeing\nthe relative momentum of interacting atoms), while the\neven-wave repulsion is in\fnitely strong, then the prob-\nlem is exactly solvable and can be mapped onto two-\ncomponent bosons with SU(2) spin rotation symmetry.\nFor the latter case the ground state is known to be ferro-\nmagnetic [14{16]. However, the spin rotation symmetry\nbreaks if the odd-wave interaction is resonant, since it\nthen depends on the spin projections of colliding par-\nticles. Therefore, in this regime the exact solution is\nno longer available, and in order to make conclusions\nabout the character of the ground state we have to em-\nploy many-body perturbation theory.\nWe consider a 1D two-component Fermi gas in free\nspace and assume that the intercomponent contact (even-\nwave) interaction is in\fnitely repulsive. Since it takes\nplace between two particles with zero total (pseudo)spin,\nit is only present in the non-ferromagnetic phases. There-\nfore, omitting the odd-wave interaction, the ferromag-\nnetic phase represents an ideal single-component Fermi\ngas, with the Fermi momentum kF=\u0019n, and the total\nenergy is equal to the kinetic energy:\nEf=Ekin=\u00192~2n2\n6mN=1\n3EFN; (1)\nwherenis the 1D density, Nis the total number of par-\nticles,mis the mass of a particle, and EF=~2k2\nF=2mis\nthe Fermi energy. The non-ferromagnetic phases in this\ncase are described by the exactly solvable Yang-Gaudin\nmodel [17, 18], and for any \fnite contact repulsion they\nhave a lower energy than the ferromagnetic phase. If the\nrepulsion is in\fnite, then the wavefunction vanishes when\ntwo particles approach each other at zero distance, and\nall spin con\fgurations are degenerate with the energy\nequal toEkin[17{19].\nThe odd-wave interaction can be both inter- and in-\ntracomponent. However, we are interested in the regime\nwhere this interaction is resonant. Since for the most\nimportant case of40K atoms the resonance in the odd-\nwave channel is present only between two atoms in the\n9=2;\u00007=2 states we con\fne ourselves to the odd-wave in-\nteraction between these states. Below the state 9 =2;\u00007=2\nis denoted as spin- \", and the state 9 =2;\u00009=2 as spin-#.\nMoreover, we assume that although the odd-wave inter-\naction is resonant, it is not too strong (a more precise\ncondition will be given later), and still can be treated as\nperturbation. The fact that one can use a perturbative\napproach in a 1D odd-wave interacting system (in con-\ntrast to the even-wave interaction) \fnds its origin in the\nabsence of a weakly bound state in a su\u000eciently shallow\nattractive potential. For the ferromagnetic many-body\nsystem our perturbative results perfectly agree with theexisting Bethe Ansatz solution [20].\nThus, to zero order the kinetic energy Ekinis the same\nin any spin con\fguration (as a consequence of the in\f-\nnite even repulsion) and gives the main contribution to\nthe total energy Eof the system, while the odd-wave in-\nteraction provides a small correction, which we derive up\nto the second order in perturbation theory. For the non-\nferromagnetic phases we employ the single-component\nmomentum distribution functions N\"(k) andN#(k) that\nwe obtain by solving numerically the Bethe Ansatz equa-\ntions for the Yang-Gaudin model at an in\fnite intercom-\nponent repulsion (see Supplemental Material). Consid-\nering equally populated \"and#internal states, N\"(k) =\nN#(k) =N(k), we have\n\u0002+1\n\u00001dk\n2\u0019N(k) =n\n2; (2)\n\u0002+1\n\u00001Ldk\n2\u0019~2k2\n2mN(k) =Ekin=2 =\u00192~2n2\n12mN; (3)\nwithLbeing the size of the system. For the ferromag-\nnetic phase we use the Fermi step momentum distribu-\ntionN(k) =\u0012(kF\u0000jkj)=2. The kinetic energy obtained\nby using the calculated non-ferromagnetic distributions\ndirectly in Eq. (3) di\u000bers from Ekinby less than 0 :3% in\nthe antiferromagnetic phase and by approximately 0 :5%\nin the paramagnetic phase.\nIn order to develop many-body perturbation theory we\nfollow the method used in Refs. [21, 22]. We de\fne the\no\u000b-shell scattering amplitude\nf(k0;k) =\u00021\n\u00001dxe\u0000ik0xV(x) k(x); (4)\nwhereV(x) is the interaction potential, k(x) is the true\nwavefunction of the relative motion with momentum k=\n(k1\u0000k2)=2, withk1,k2andk0\n1,k0\n2being the particle mo-\nmenta in the incoming and outgoing scattering channels.\nForjk0j=jk0\n1\u0000k0\n2j=2 =jkjwe have the on-shell ampli-\ntude. Then, the total energy is E=Ekin+~E(1)+~E(2),\nwhere the \frst- and second-order corrections are given by\n(see Supplemental Material):\n~E(1)=1\nLX\nk1;k2~fodd(k)N(k1)N(k2); (5)\n~E(2)=\u00001\nL2X\nk1;k2;k0\n14m\n~2~fodd(k0;k)~fodd(k;k0)\nk2\n1+k2\n2\u0000k02\n1\u0000k02\n2\n\u0002N(k1)N(k2)N(k0\n1);(6)\nwithk1+k2=k0\n1+k0\n2. The amplitude ~foddis di\u000ber-\nent from the odd-wave part of (4) by the absence of the3\nimaginary term in the denominator (see Supplemental\nMaterial). The terms ~E(1)and ~E(2)are the two-body\n(mean-\feld) and the many-body, or beyond mean-\feld,\ncontributions to the interaction energy. As the 1D regime\nis obtained by tightly con\fning the motion of particles in\ntwo directions to zero point oscillations, the odd-wave o\u000b-\nshell scattering amplitude in the vicinity of the resonance\nis given by (see [23] and Supplemental Material):\n~fodd(k0;k) =2~2\nmk0klp\n1 +\u0018plpk2; (7)\nwhere the parameters lpand\u0018pof the 1D odd-wave scat-\ntering can be expressed through the parameters of the\n3Dp-wave scattering as\nlp= 3a?\u0014a3\n?\nw1+A\u0015\u00001\n; \u0018p=\u000b1a2\n?\n3: (8)\nHerew1and\u000b1are the 3D scattering volume and e\u000bec-\ntive range, respectively, a?=p\n~=(m!?) is the exten-\nsion of the wavefunction in the directions tightly (har-\nmonically) con\fned with frequency !?, and the numeri-\ncal constant isA=\u00003p\n2\u0010(\u00001=2)\u00190:88, with\u0010(\u00001=2)\nbeing the Riemann zeta function. For40K atoms near\nthep-wave Feshbach resonance the magnetic \feld depen-\ndence ofw1and\u000b1has been measured in the JILA ex-\nperiments [24]. Near the resonance in 3D the scattering\nvolumew1changes from in\fnitely negative to in\fnitely\npositive, whereas the e\u000bective range \u000b1remains practi-\ncally constant and equal to 4 \u0002106cm\u00001. On the positive\nside of the resonance ( w1>0) in 3D one has the forma-\ntion of rapidly decaying p-wave molecules [25{29], and\na similar phenomenon is expected in 1D. The issue of\ninelastic losses is discussed in more detail below, but in\nwhat follows we consider only the case of attractive odd-\nwave interaction, i.e. lp<0. The energy corrections (5)\nand (6) can be rewritten as (see Supplemental Material)\n~E(1)=\u0000Ekin\u001a1\n2\u0019\u0011+3\n16\u0019\u0014\u00112I(Q)\u001b\n; (9)\n~E(2)=3\n4\u00192\u00112J(Q)Ekin; (10)\nwhere\u0011=kFjlpj,\u0014=kF\u0018p,Q=\u0011\u0014, and we took into\naccount that in any spin con\fguration the momentum\ndistribution is a universal function of k=kF. The inte-\ngralsI(Q) andJ(Q) are given by\nI(Q) = +1\n\u00001dx1dx2N(x1)N(x2)(x1\u0000x2)4\n1\u0000Q\n4(x1\u0000x2)2;(11)\nJ(Q) = +1\n\u00001dx1dx2dx3N(x1)N(x2)N(x3)\nx1\u0000x3\n\u0002(x1\u0000x2)2\n\u0010\n1\u0000Q\n4(x1\u0000x2)2\u0011(x1+x2\u00002x3)2\n\u0010\n1\u0000Q\n4(x1+x2\u00002x3)2\u0011;(12)withxi=ki=kFbeing a dimensionless momentum, and\nthe symbol\u001f\ndenoting the principal value of the inte-\ngral. The choice of a particular spin con\fguration is\nencoded in the momentum distribution N(k=kF), and\nfrom Eqs. (9)-(12) it is evident that the odd-wave in-\nteraction splits the energies of di\u000berent phases only if\n\u00146= 0. The unperturbed momentum distributions for the\nferro-, antiferro-, and paramagnetic states are displayed\nin Fig.1. At k\u001dkFthe momentum distributions in the\nnon-ferromagnetic states behave as N(k)!C=k4, where\nCis Tan's contact [30, 31]. In the aniferromagnetic phase\nits value isC=k4\nF= 2 ln(2)=3\u00192\u00190:047 [31], and in the\nparamagnetic phase we obtain C=k4\nF\u00190:016.\nFIG. 1. Momentum distributions in the ferro-, antiferro-, and\nparamagnetic phases. The latter two are for the Yang-Gaudin\nmodel at an in\fnite repulsion.\nIn the limit of Q!0 we haveI= 2(1 + 3D)=3\nandJ= 1=2, whereD=\u00011\n\u00001dx\u0002\nN(x)x4\u0000C=k4\nF\u0003\n. For\na \fniteQwe calculate the integrals (11) and (12) numer-\nically.\nRealization of the 1D regime requires the Fermi energy\nto be much smaller than the tight con\fnement frequency:\nEF=~2k2\nF\n2m\u001c~!?: (13)\nFor realistic con\fnement frequencies !?in the range\nfrom 50 to 150 kHz, the condition (13) requires the Fermi\nmomentum kF.105cm\u00001(which corresponds to densi-\ntiesn.3\u0002104cm\u00001andEF.1\u0016K). The con\fnement\nlengtha?for such frequencies is from 400 to 700 \u0017A. Then,\ntaking the potassium value 4 \u0002106cm\u00001for the e\u000bective\nrange\u000b1and using relations (8) we see that the param-\neter\u0014ranges from 1 to 5. In the perturbative regime we\nrequire\u0011=\u0019=njlpj\u001c1, i.e. one should not be too close\nto the resonance, and in order to stay within the limits\nof perturbation theory we put \u0011<0:8.\nWe then calculate the total energy of the gas up\nto the second order in perturbation theory for the\nferro-, antiferro-, and paramagnetic phases. The results\nare presented in Fig.2, which shows that the di\u000berence4\nbetween the energies of the ferro- and non-ferromagnetic\nstates is the largest at \u0011\u00181=\u0014and decreases signi\fcantly\nas\u0014grows.\nFor a gas of40K atoms with a density n\u00193\u0002\n104cm\u00001(EF\u0019540 nK) under the transverse con-\n\fnement with frequency !?\u0019100 kHz we have\n\u0014=\u0019n\u000b 1~=(3m!?)\u00193:1, and at \felds slightly lower\nthan 199 G the parameter \u0011is approximately 0 :36. Then\nthe ferromagnetic state has the lowest energy, and the\nenergies of antiferro- and paramagnetic states are close\nto each other. The energy di\u000berence ( Ep\u0000Ef)=N\nis about 0:03EFor 16 nK. Increasing the con\fnement\nstrength to !?\u0019120 kHz, we obtain \u0014\u00192:6, and\nwith\u0011\u00190:43 at the magnetic \feld B\u0019199 G the energy\ndi\u000berence ( Ep\u0000Ef)=Nbecomes approximately 20 nK.\nThus, the ferromagnetic state can be observed at tem-\nperatures below 20 nK.\nThe regimes described above ensure that the even re-\npulsion is in\fnitely strong, although at \frst sight the\ncorresponding magnetic \felds are not close enough to\nthes-wave resonance. However, in the 1D geometry ob-\ntained by tightly con\fning particles in two directions, the\ncoupling constant for the even contact interaction is [32]\ng1D=2~2a\nma?(a?\u00001:03a): (14)\nFor the con\fning frequencies of 100 and 150 kHz the\nharmonic length a?is about 500 and 400 \u0017A, respec-\ntively. In a \feld close to 199 Gauss the scattering length\nisa\u0019400\u0017A. Thus, due to the con\fnement-induced\nresonance, one can achieve an in\fnite contact repulsion\ng1D!1 .\nIt is well known that in three dimensions p-wave Fesh-\nbach resonances are su\u000bering of rapid inelastic losses [25{\n29]. There are two types of inelastic collisional processes.\nThe \frst one is three-body recombination, which is espe-\ncially pronounced if there are weakly bound dimer states.\nHowever, weakly bound dimer p-wave states are expected\nonly on the positive side of the resonance ( lp>0), and\non the negative side ( lp<0 and attractive interactions)\nthe three-body recombination should not be very danger-\nous, at least slightly away from the resonance. Another\ndecay process is two-body relaxation. The internal state\n9=2;\u00007=2 has a higher energy than 9 =2;\u00009=2. Therefore,\nthe state 9=2;\u00007=2 can undergo collisional relaxation to\nthe 9=2;\u00009=2 state. In 3D the rate of this process is\nespecially high very near the resonance [25, 33].\nIn \felds slightly higher than 199 G the mea-\nsured rate constant of three-body recombination in 3D\nis\u000b3D\nrec\u001810\u000025cm6/s and the rate constant of two-body\nrelaxation is \u000b3D\nrel\u001810\u000014cm3/s [25]. The measure-\nments were done in a pure gas of 9 =2;\u00007=2 atoms, and\nin a mixture of 9 =2;\u00007=2 and 9=2;\u00009=2 states it can be\nsomewhat higher. The temperature in the experiment\nwas from 1 \u0016K to 3\u0016K, so that one expects about the\nFIG. 2. The total energy (in units of Ekin=EFN=3) versus\u0011\nin the ferro-, antiferro- and paramagnetic state for \u0014= 3,\n\u0014= 4, and\u0014= 5.\nsame rate constants at EF\u00181\u0016K and much lower tem-\nperatures. In order to transform these results to 1D one\nshould recall that the inelastic processes occur at atomic\ninterparticle distances. We thus may integrate out the\nmotion in the tightly con\fned directions [34]. This leads\nto\u000b1D\nrel\u0019\u000b3D\nrel=2\u0019a2\n?[35]. With the above speci\fed \u000b3D\nrel\nand densities n\u0018104or 3\u0002104cm\u00001we obtain a relax-\nation time of about a second.\nRegarding the three-body recombination the situation\nis more peculiar. For the collision of three particles in one\nand the same internal state in 1D one has an extra sup-\npression by a factor of EF=E\u0003compared to 3D [36]. The\nquantityE\u0003is a typical energy in the molecular problem,\nand one has E\u0003\u0018~2=mR2\ne, whereRe\u001850\u0017A is the ra-5\ndius of interaction between particles. So, E\u0003\u00181 mK or\neven larger and there is an extra suppression by 3 orders\nof magnitude for EF\u00181\u0016K. Integrating out the par-\nticle motion in the tightly con\fned direction we obtain\n\u000b1D\nrec\u0018\u000b3D\nrec(EF=E\u0003)=3\u00192a4\n?[35]. Again, for the above\nmentioned parameters we obtain the decay time about a\nsecond at 1D densities in between 104and 3\u0002104cm\u00001.\nThe ferromagnetic state can be viewed as a composition\nof identical fermions, and hence this estimate remains\nvalid in this phase.\nWe thus see that in 1D the issue of inelastic decay\nprocesses is not as crucial as in 3D. In this respect the\nsituation is somewhat similar to the one in recent exper-\niments with strongly interacting bosons [37, 38].\nIn conclusion, we showed that there is a realistic pos-\nsibility to \fnd itinerant ferromagnetic states in 1D two-\ncomponent Fermi gases, and a promising system is the\ngas of40K atoms. This will require \fne tuning of the in-\nteraction between particles by varying the magnetic \feld\nand the strength of the tight con\fnement. The required\ntemperatures are about 10 to 20 nanokelvin, which is\nachievable with present facilities.\nWe would like to thank V. Gritsev, D. Petrov, and\nM. Zvonarev for useful discussions. We acknowledge\nsupport from IFRAF and from the Dutch Foundation\nFOM. The research leading to these results has received\nfunding from the European Research Council under Eu-\nropean Community's Seventh Framework Programme\n(FR7/2007-2013 Grant Agreement no. 341197). This\nwork has been partially supported by the NNSFC under\ngrant numbers 11374331 and 11304357.\n[1] H. von L ohneysen, A. Rosch, M. Vojta, and P. W ol\re,\nRev. Mod. Phys. 79, 1015 (2007).\n[2] E.C. Stoner, Phil. Mag. 15, 1018 (1933).\n[3] D.C. Mattis, The Theory of Magnetism (Harper & Row,\nNew York, 1965).\n[4] T. Sogo and H. Yabu, Phys. Rev. A 66, 043611 (2002);\nR. A. Duine and A. H. MacDonald, Phys. Rev. Lett. 95,\n230403 (2005); H. Zhai, Phys. Rev. A 80, 051605 (2009);\nG.J. Conduit and B.D. Simons, Phys. Rev. A 79, 053606\n(2009), Phys. Rev. Lett. 103, 200403 (2009); G.J. Con-\nduit, A.G. Green, and B.D. Simons, Phys. Rev. Lett. 103,\n207201 (2009); L. J. LeBlanc, J. H. Thywissen, A. A.\nBurkov, and A. Paramekanti, Phys. Rev. A 80, 013607\n(2009); I. Berdnikov, P. Coleman, and S. H. Simon, Phys.\nRev. B 79, 224403 (2009).\n[5] S. Pilati, G. Bertaina, S. Giorgini, and M. Troyer, Phys.\nRev. Lett. 105, 030405 (2010).\n[6] S.-Y. Chang, M. Randeria, and N.Trivedi, PNAS 108, 51\n(2011).\n[7] G.-B. Jo, Y.-R. Lee, J.-H. Choi, C.A. Christensen, T.H.\nKim, J.H. Thywissen, D.E. Pritchard, and W. Ketterle,\nScience 325, 1521 (2009).\n[8] C. Sanner, E.J. Su, W. Huang, A. Keshet, J. Gillen, and\nW. Ketterle, Phys. Rev. Lett. 108, 240404 (2012).[9] D. Pekker, M. Babadi, R. Sensarma, N. Zinner, L. Pol-\nlet, M.W. Zwierler, and E. Demler, Phys. Rev. Lett. 106,\n050402 (2011).\n[10] Proposals to obtain the ferromagnetic and spin segre-\ngated states by sweeping across a Feshbach resonance\nfrom strongly repulsive to attractive interaction in a two-\ncomponent harmonically trapped 1D Fermi gas were made\nin S.E. Gharashi and D. Blume, Phys. Rev. Lett. 11,\n045302 (2013); X. Cui and T.-L. Ho, Phys. Rev. A 89,\n023611 (2014).\n[11] E.H. Lieb and D.C. Mattis, Phys. Rev. 125, 164 (1962).\n[12] C.A. Regal, C. Ticknor, J.L. Bohn, and D.S. Jin, Phys.\nRev. Lett. 90, 053201 (2003).\n[13] T. Loftus, C.A. Regal, C. Ticknor, J.L. Bohn, and D.S.\nJin, Phys. Rev. Lett. 88, 173201 (2002).\n[14] Y.-Q. Li, S.-J. Gu, Z.-J. Ying, and U. Eckern, Europhys.\nLett. 61, 368 (2003).\n[15] M.T. Batchelor, M. Bortz, X.-W. Guan, and N. Oelkers,\nJ. Stat. Mech. 2006 , P03016 (2003).\n[16] X.-W. Guan, M.T. Batchelor, and M. Takahashi, Phys.\nRev. A 76, 043617 (2007).\n[17] M. Gaudin, Phys. Lett. A 24, 55 (1967).\n[18] C.-N.Yang, Phys. Rev. Lett. 19, 1312 (1967).\n[19] X.-W. Guan, M.T. Batchelor, C. Lee, Rev. Mod. Phys.\n80, 1633 (2013).\n[20] A. Imambekov, A.A. Lukyanov, L.I. Glazman, and V.\nGritsev, Phys. Rev. Lett. 104, 040402 (2010).\n[21] A.A. Abrikosov, I.M. Khalatnikov, Sov. Phys. JETP 6,\n888 (1958).\n[22] Z.-K. Lu, G.V. Shlyapnikov, Phys. Rev. A 85, 023614\n(2012).\n[23] L. Pricoupenko, Phys. Rev. Lett. 100, 170404 (2008).\n[24] C. Ticknor, C.A. Regal, D.S. Jin, and J.L. Bohn, Phys.\nRev. A 69, 042712 (2004).\n[25] C.A. Regal, C. Ticknor, J.L. Bohn, and D.S. Jin, Phys.\nRev. Lett. 90, 053201 (2003).\n[26] F. Chevy, E.G.M. van Kempen, T. Bourdel, J. Zhang,\nL. Khaykovich, M. Teichmann, L. Tarruell, S.J.J.M.F.\nKokkelmans, and C. Salomon, Phys. Rev. A 71, 062710\n(2005).\n[27] Y. Inada, M. Horikoshi, S. Nakajima, M. Kuwata-\nGonokami, M. Ueda, and T. Mukaiyama, Phys. Rev. Lett.\n101, 100401 (2008).\n[28] P. Zhang, P. Naidon, and M. Ueda, Phys. Rev. A 82,\n062712 (2010).\n[29] J. Fuchs, C. Ticknor, P. Dyke, G. Veeravalli, E. Kuhnle,\nW. Rowlands, P. Hannaford, and C. J. Vale, Phys. Rev.\nA77, 053616 (2008).\n[30] S. Tan, Ann. Phys. (N.Y.), 323, 2952 (2008).\n[31] M. Barth, W. Zwerger, Ann. Phys. (N.Y.), 326, 2544\n(2011).\n[32] M. Olshanii, Phys. Rev. Lett. 81, 938 (1998).\n[33] J. Bohn (private communication).\n[34] D.M. Gangardt and G.V. Shlyapnikov, Phys. Rev. Lett.\n90, 010401 (2003).\n[35] A detailed study of inelastic decay processes for our sys-\ntem will be given elsewhere.\n[36] N.P. Mehta, B.D. Esry, and C.H. Greene, Phys. Rev. A\n76, 022711 (2007).\n[37] E. Haller, M. Gustavsson, M.J. Mark, J.G. Danzl, R.\nHart, G. Pupillo, H.-C. N agerl, Science 352, 5945 (2009).\n[38] E. Haller, M.J. Mark, R. Hart, J.G. Danzl, L. Re-\nichs ollner, V. Melezhik, P. Schmelcher, and H.-C. N agerl,\nPhys. Rev. Lett. 104, 153203 (2010); E. Haller, M. Rabie,1\nM. J. Mark, J. G. Danzl, R. Hart, K. Lauber, G. Pupillo,\nand H.-C. N agerl, Phys. Rev. Lett. 107, 230404 (2011).\nSupplemental Material: Itinerant ferromagnetism in 1D two-component Fermi gases\nIn the Supplemental Material we \frst discuss the scattering of identical fermions in 1D, obtain the expression\nfor the odd-wave o\u000b-shell scattering amplitude, and derive two-body and many-body contributions to the odd-wave\ninteraction energy of a 1D Fermi gas. We then describe our calculation of the momentum distribution in the antiferro-\nand paramagnetic states.\nScattering amplitude in 1D. The o\u000b-shell scattering amplitude, de\fned by Eq. (4) of the main text, is the sum of\nthe even-wave and odd-wave partial amplitudes:\nf(k0;k) =feven(k0;k) +fodd(k0;k); (S.1)\nwhere\nfeven(k0;k) =\u00021\n\u00001dx0cosk0x0V(x0) even(k;x0); fodd(k0;k) =\u0000i\u00021\n\u00001dx0sink0x0V(x0) odd(k;x0); (S.2)\nand even, oddare the partial wavefunctions of the relative motion in the even-wave and the odd-wave channels,\nrespectively. For jk0j=jkjone has the on-shell amplitudes, which follow from from Eq. (S.2) putting k0=k.\nThese amplitudes enter the asymptotic expression for the wavefunction of the relative motion at interparticle separa-\ntionsjxj!1 [S1]:\n k(x) =eikx\u0000im\n2~2keikjxj\u0000\nfeven(k) + sign(x)fodd(k)\u0001\n: (S.3)\nIn the quasi1D regime obtained by tightly con\fning the motion of particles in two directions to zero point oscillations,\nthe on-shell scattering amplitude in the odd-wave channel was calculated in Ref. [S2], and it reads:\nfodd(k) =2~2\nmk2\n1=lp+ik+\u0018pk2; (S.4)\nwhere parameters lpand\u0018pare given by Eq. (8) of the main text.\nWe now proceed with the derivation of the odd-wave o\u000b-shell scattering amplitude. The asymptotic form of the\nwavefunction odd(k;x) atjxj!1 can be written as\n odd(k;x)/sin\u0000\nkjxj\u0000\u000eodd(k)\u0001\n; (S.5)\nwhere\u000eodd(k) is the scattering phase shift. The relation between the odd-wave phase shift and the corresponding\nscattering amplitude follows from Eq. (S.5) and the odd-wave part of Eq. (S.3):\nfodd(k) =2~2k\nmtan\u000eodd(k)\n1 +itan\u000eodd(k); (S.6)\nwith tan\u000eodd(k) =lpk=\u0000\n1 +lp\u0018pk2\u0001\n. Let us rewrite the wavefunction of the odd-wave channel odd(k;x) in the\nfollowing form:\n odd(k;x) =i~ odd(k;x)\n1 +itan\u000eodd(k); (S.7)\nwhere ~ odd(k;x) = sinkx\u0000sign(x) tan\u000eodd(k) coskxis real. Then, the odd-wave o\u000b-shell scattering amplitude can\nbe represented as\nfodd(k0;k) =~fodd(k0;k)\n1 +itan\u000eodd(k); (S.8)2\nwhere ~fodd(k0;k) is also real:\n~fodd(k0;k) =\u00021\n\u00001dx0sink0x0V(x0)~ odd(k;x0): (S.9)\nAssuming that the phase shift is small, one obtains the following relation:\nfodd(k0;k) =~fodd(k0;k)\u0000im\n2~2k~f2\nodd(k0;k): (S.10)\nSubstituting k0=kinto Eqs. (S.8)-(S.10) we obtain similar relations for the on-shell amplitudes fodd(k) and ~fodd(k).\nIn the case of low-energy scattering we may put sin k0x0\u0019k0x0and sinkx0\u0019kx0in the expressions for ~fodd(k0;k) and\n~fodd(k), which shows that in the odd-wave channel one obtains the o\u000b-shell amplitude from the on-shell amplitude\nby simply replacing kin (S.6) with k0. Then, using Eqs. (S.6) and (S.8) we arrive at Eq. (7) of the main text for the\nodd-wave o\u000b-shell scattering amplitude ~fodd(k0;k).\nTwo-body and many-body contributions to the interaction energy. An in\fnitely strong even-wave contact repulsion\ncan be transferred to the boundary condition for the wavefunction. The interaction part of the Hamiltonian then\ncontains only the odd-wave interaction, which in our case is present solely between \"-state particles:\n^Hint=1\n2LX\nk1;k2;qV(q) ^ay\nk1+q^ay\nk2\u0000q^ak2^ak1; (S.11)\nwhere ^ay\nk;^akare the creation and annihilation operators of \"-fermions, and V(q) is the Fourier transform of the\ninteraction potential in the odd-wave channel:\nV(q) =\u00021\n\u00001dxV(x)e\u0000iqx=\u00021\n\u00001dxV(x) cosqx; (S.12)\nwhere we took into account that V(x) is even. Then, the \frst-order correction is given by the diagonal matrix element\nof^Hint:\nE(1)=1\n2LX\nk1;k2[V(0)\u0000V(k2\u0000k1)]N(k1)N(k2): (S.13)\nThe second-order correction to the energy of a state jjiis given by\nE(2)\nj=X\nm6=jjVjmj2\nEj\u0000Em; (S.14)\nwhere the summation is over the eigenstates of the non-interacting system, and the non-diagonal matrix element\nVjm=hmj^Hintjjiis related to the scattering of two particles from the initial state k1;k2to an intermediate state k0\n1;k0\n2.\nIn our case, the symbol jcorresponds to the ground state, and the symbol mto excited states. Then, taking into\naccount the momentum conservation law k1+k2=k0\n1+k0\n2, we obtain the following expression for the quantity jVjmj2:\njVjmj2=1\n(2L)2X\nk1;k2jV(k0\n1\u0000k1)\u0000V(k0\n2\u0000k1)j2N(k1)N(k2)\u0000\n1\u0000N(k0\n1)\u0001\u0000\n1\u0000N(k0\n2)\u0001\n; (S.15)\nand the second-order correction becomes:\nE(2)=1\n(2L)2X\nk1;k2;k0\n1jV(k0\n1\u0000k1)\u0000V(k0\n2\u0000k1)j2\n~2(k2\n1+k2\n2\u0000k02\n1\u0000k02\n2)=2mN(k1)N(k2)\u0000\n1\u0000N(k0\n1)\u0001\u0000\n1\u0000N(k0\n2)\u0001\n: (S.16)\nIt is evident that the second-order correction diverges at large k0\n1because of the term proportional to N(k1)N(k2).\nThis arti\fcial divergence can be eliminated if one expresses E(1)andE(2)in terms of a real physical quantity |\nthe scattering amplitude. The relation between the Fourier component of the interaction potential and the o\u000b-shell\nscattering amplitude is given by [S1]\nf(k0;k) =V(k0\u0000k) +1\nLX\nk00V(k0\u0000k00)f(k00;k)\nEk\u0000Ek00+i0; (S.17)3\nwhereEkandEk00are relative collision energies, and we have Ek\u0000Ek00=~2(k2\n1+k2\n2\u0000k002\n1\u0000k002\n2)=2m, withk1;k2(k00\n1;k00\n2)\nbeing the momenta of colliding particles in the initial (intermediate) state. Equation (S.17) allows us to rewrite the\n\frst-order correction as\nE(1)=1\n2LX\nk1;k2[f(k;k)\u0000f(\u0000k;k)]N(k1)N(k2)\u00001\n2L2X\nk1;k2;k0[V(k\u0000k0)\u0000V(\u0000k\u0000k0)]f(k0;k)\n~2(k2\u0000k02+i0)=mN(k1)N(k2):(S.18)\nIn the second term of Eq. (S.18) we represent f(k0;k) according to Eq. (S.1) and keep only the odd-wave\npartfodd(k0;k), because the quantity [ V(k\u0000k0)\u0000V(\u0000k\u0000k0)] is odd in k0, and the terms containing feven(k0;k)\nvanish after the integration over dk0. Then, we replace the Fourier components by the exact scattering amplitudes,\nwhich gives [ V(k\u0000k0)\u0000V(\u0000k\u0000k0)] = [f(k;k0)\u0000f(\u0000k;k0)] = 2fodd(k;k0), and the \frst-order correction becomes:\nE(1)=1\nLX\nk1;k2fodd(k)N(k1)N(k2)\u00001\nLX\nk1;k2N(k1)N(k2)\u00021\n\u00001dk0\n2\u0019m\n~2fodd(k0;k)fodd(k;k0)\nk2\u0000k02+i0: (S.19)\nThe contribution of the pole at k0=kto the integral in Eq. (S.19) gives \u0000imf2\nodd(k)=2~2kfor each term in the sum\noverk1,k2, and we can use here ~fodd(k) instead of fodd(k). At the same time, the second term of the right hand side\nof Eq. (S.10), being substituted into the \frst term of Eq. (S.19), cancels the contribution of the pole in the second\nterm of Eq. (S.19). Therefore, we use the amplitude ~fodd(k) in the \frst term of Eq. (S.19) and take the principal\nvalue of the integral in the second term. This leads to the following expression for the \frst-order correction:\nE(1)=1\nLX\nk1;k2~fodd(k)N(k1)N(k2)\u00001\nL2X\nk1;k2;k0\n12m\n~2~fodd(k0;k)~fodd(k;k0)\nk2\n1+k2\n2\u0000k02\n1\u0000k02\n2N(k1)N(k2): (S.20)\nThe second-order correction can also be expressed in terms of the scattering amplitude. Taking into ac-\ncount that V(k0\n1\u0000k1) =V(k0\u0000k) andV(k0\n2\u0000k1) =V(\u0000k\u0000k0), we writejV(k0\n1\u0000k1)\u0000V(k0\n2\u0000k1)j2=\n[V(k\u0000k0)\u0000V(k+k0)] [V(k0\u0000k)\u0000V(\u0000k\u0000k0)], where we used the fact that Fourier components V(q) are real\nand even functions. Then for the second-order correction we obtain:\nE(2)=1\nL2X\nk1;k2;k0\n12m\n~2fodd(k0;k)fodd(k;k0)\nk2\n1+k2\n2\u0000k02\n1\u0000k02\n2N(k1)N(k2)\u0000\n1\u0000N(k0\n1)\u0001\u0000\n1\u0000N(k0\n2)\u0001\n; (S.21)\nwhere we may use the amplitudes ~fodd(k0;k) and ~fodd(k;k0) because the contribution of tan \u000eodd(k) in the denominator\nof Eq. (S.8) is negligible. Then, the divergent term proportional to N(k1)N(k2) in Eq. (S.21) and the (divergent)\nsecond term of Eq. (S.20) exactly cancel each other. Note that in Eq. (S.21) the term proportional to the product of\nfour occupation numbers vanishes, since its numerator is symmetrical and the denominator is antisymmetrical with\nrespect to an interchange of k1;k2andk0\n1;k0\n2. Two terms containing the product of three occupation numbers are\nequal to each other, because the expression (S.21) for E(2)is symmetrical with respect to an interchange of k0\n1and\nk0\n2. Therefore, the sum of the \frst- and second-order corrections can be written as E(1)+E(2)=~E(1)+~E(2), where\n~E(1)and ~E(2)are given by\n~E(1)=1\nLX\nk1;k2~fodd(k)N(k1)N(k2) =2~2\nmLX\nk1;k2lpk2\n1 +lp\u0018pk2N(k1)N(k2); (S.22)\n~E(2)=\u00001\nL2X\nk1;k2;k0\n14m\n~2~fodd(k0;k)~fodd(k;k0)\nk2\n1+k2\n2\u0000k02\n1\u0000k02\n2N(k1)N(k2)N(k0\n1)\n=\u000016~2\nmL2X\nk1;k2;k0\n1l2\npk02k2\nk2\n1+k2\n2\u0000k02\n1\u0000k02\n2N(k1)N(k2)N(k0\n1)\n(1 +lp\u0018pk02)(1 +lp\u0018pk2):(S.23)\nWe then reduce Eqs. (S22) and (S23) to equations (9)-(12) of the main text.\nMomentum distributions for the antiferro- and paramagnetic states. The momentum distribution functions N\"(k)\nandN#(k) for the antiferro- and paramagnetic states can be calculated from the Bethe Ansatz wave functions of the4\nYang-Gaudin model with an in\fnite repulsion. To this end, we will use the method proposed by Ogata and Shiba [S3].\nThe one-dimensional two-component Fermi gas with contact interactions is described by the Hamiltonian [S4, S5]\n^H=\u0000PN\ni=1@2=@x2\ni+2cP\ni 0\nwhereas\u0012(x) = 0 forx < 0. In the above equations, we denoted x=fx1;x2;\u0001\u0001\u0001;xNg,\u001b=f\u001b1;\u001b2;\u0001\u0001\u0001;\u001bNgand\nA=A\u001b1\u0001\u0001\u0001\u001bN(Q;P) are superposition coe\u000ecients. Here \u001bj=\u00061=2 stand for the spin projection of the j-th particle.\nFor periodic boundary conditions the wave numbers fkjgwithj= 1;2:::;N are subject to the following Bethe\nAnsatz equations (BAE):\neikjL=MY\ni=1kj\u0000\u0016i+ ic=2\nkj\u0000\u0016i\u0000ic=2; (S.25)\nNY\nj=1\u0016i\u0000kj\u0000ic=2\n\u0016i\u0000kj+ ic=2=\u0000MY\ni0=1\u0016i\u0000\u0016i0\u0000ic\n\u0016i\u0000\u0016i0+ ic:\nHereMis the number of down-spin fermions, and \u0016iwithi= 1;:::;M are spin rapidities.\nWe observe that the BAE (S.25) decouple into two parts in terms of charge and spin degrees of freedom as c!1 .\nThe second set of equations in the BAE (S.25) reduces to the BAE for the spin-1 =2 Heisenberg XXX model with the\nscaling\u0016\u000b!\u0016\u000bc. In this limit, the wave function (S.24) can be simpli\fed as\n\t\u001b(x) =X\nQP(\u00001)Q+P\b(y1;:::;yM) eiP\njkPjxQj: (S.26)\nHere\b(y1;:::;yM) is the eigenstate of the spin XXX model with Mdown-spins in the N-site lattice [S3, S5].\nUsing the wave function (S.26), we \frst calculate the momentum distribution in an analytical fashion. Then the \fnal\nmomentum distributions can be obtained by further numerical calculation. The momentum distribution is de\fned as\nN\u001b(k) =h^ y\n\u001b(k)^ \u001b(k)i, which is the Fourier transform of the density matrix N\u001b(k) =\u0003\ndydy0N\u001b(y0;y)eik(y\u0000y0). The\ndensity matrix is given by\nN\u001b(y0;y) =h y\n\u001b(y0) \u001b(y)i=N!\u0002x3\n0dx2\u0002x4\n0dx3\u0001\u0001\u0001\u0002xN\n0dxN\u00001\u0002L\n0dxNX\n\u001b2\u0001\u0001\u0001\u001bN\t\u0003\n\u001b(x(l))\t\u001b(x(r));(S.27)\nwhere we used the following notations:\nx(l)=fy0;x2;x3;\u0001\u0001\u0001;xNg;x(r)=fy;x2;x3;\u0001\u0001\u0001;xNg;\u001b=f\u001b;\u001b;\u001b 3;\u0001\u0001\u0001;\u001bNg: (S.28)\nIn the limit c!1 , the superposition coe\u000ecients satisfy the relations [S3]:\n~A(Q(ab);P(ab)) =^Pab~A(Q(ba);P(ba));~A(Q(ab);P(ab)) =\u0000~A(Q(ab);P(ba));\nwhere\nQ(ab)=fQ1;Q2;\u0001\u0001\u0001;Qa;Qb;\u0001\u0001\u0001;QNg;Q(ba)=fQ1;Q2;\u0001\u0001\u0001;Qb;Qa;\u0001\u0001\u0001;QNg:\nSubstituting the wave function (S.26) into (S.27), the density matrix is thus rewritten as\nN\u001b(y0;y) =N!\u0002x3\n0dx2\u0002x4\n0dx3\u0001\u0001\u0001\u0002L\n0dxNX\nP;P0(\u00001)P0+P+Q+Q0e\u0000iPN\nj=1kP0\njx(l)\nQ0eiPN\nj=1kPjx(l)\nQ\n\u0002X\n\u001b2;\u0001\u0001\u0001;\u001bN\u0002\nAy\fy(Q0)\u0003\n\u001b;\u001b2;\u0001\u0001\u0001;\u001bN\u0002\n\f(Q)A\u0003\n\u001b;\u001b2;\u0001\u0001\u0001;\u001bN: (S.29)\nHere the positions yandy0in the domain x2<\u0001\u0001\u0001